Environmental Protection Technology Series
BACKWASH OF GRANULAR  FILTERS USED
              IN  WASTEWATER FILTRATION
                        Municipal Environmental Research Laboratory
                             Office of Research and Development
                            U.S. Environmental Protection Agency
                                    Cincinnati, Ohio 45268

-------
                 RESEARCH  REPORTING SERIES

 Research reports of the Office of Research and Development, U.S. Environmental
 Protection Agency, have been grouped into nine series. These nine broad cate-
 gories were established to facilitate further development and application of en-
 vironmental technology. Elimination of  traditional grouping was  consciously
 planned to foster technology transfer and a maximum interface in related fields.
 The nine series are:

       1.  Environmental Health Effects Research
      2.  Environmental Protection Technology
      3   Ecological Research
      4.  Environmental Monitoring
      5.  Socioeconomic Environmental Studies
      6,  Scientific and Technical Assessment Reports (STAR)
      7.  Interagency Energy-Environment Research and Development
      8   "Special" Reports
      9.  Miscellaneous Reports

 This report has been assigned to the ENVIRONMENTAL PROTECTION TECH-
 NOLOGY series. This series describes research perlormed to develop and dem-
 onstrate instrumentation, equipment, and methodology to repair or prevent en-
 vironmental degradation from point and non-point sources of pollution. This work
 provides the new or improved technology required for the control and treatment
 of pollution sources to  meet environmental quality standards
This document is available to the public through the National Technical Informa-
tion Service, Springfield, Virginia 22161.

-------
                                      EPA-600/2-77-016
                                      April 1977
   BACKWASH OF GRANULAR FILTERS USED IN

           WASTEWATER FILTRATION
                    by

      J. L. Cleasby and E. R. Baumann
           Iowa State University
             Ames, Iowa  50011
             Grant No. R802140
              Project Officer

               S. A. Hannah
       Wastewater Research Division
Municipal Environmental Research Laboratory
          Cincinnati, Ohio  45268
MUNICIPAL ENVIRONMENTAL RESEARCH LABORATORY
    OFFICE OF RESEARCH AND DEVELOPMENT
   U.S. ENVIRONMENTAL PROTECTION AGENCY
          CINCINNATI, OHIO  45268

-------
                          DISCLAIMER
This report has been reviewed by the Municipal Environmental Re-
search Laboratory, U.S. Environmental Protection Agency,  and ap-
proved for publication.  Approval does not signify that the con-
tents necessarily reflect the views and policies of the U.S. En-
vironmental Protection Agency, nor does mention of trade  names or
commercial products constitute endorsement or recommendation for
use.
                                ii

-------
                               FOREWORD
The Environmental Protection Agency was created because of increasing
public and government concern about the dangers of pollution to the
health and welfare of the American people.  Noxious air, foul water,
and spoiled land are tragic testimony to the deterioration of our
natural environment.  The complexity of that environment and the
interplay between its components require a concentrated and integrated
attack on the problem.

Research and development is that necessary first step in problem solu-
tion and it involves defining the problem, measuring its impact, and
searching for solutions.  The Municipal Environmental Research Laboratory
develops new and improved technology and systems for the prevention,
treatment, and management of wastewater and solid and hazardous waste
pollutant discharges from municipal and community sources, for the
preservation and treatment of public drinking water supplies, and to
minimize the adverse economic, social, health, and aesthetic effects
of pollution.  This publication is one of the products of that research;
a most vital communications link between the researcher and the user
community.

The conventional methods for treatment of municipal wastewaters fre-
quently produce effluents that will not meet local discharge require-
ments.  Granular media filters are being installed to provide tertiary
treatment for increased removals of suspended solids and particulate
BOD.  This report provides valuable information on criteria for selec-
tion of media for wastewater filters and design considerations to pro-
vide adequate cleaning of the media during backwash.
                                         Francis T. Mayo
                                         Director
                                         Municipal Environmental Research
                                         Laboratory
                                   iii

-------
                             ABSTRACT


The use of deep granular filters in waste treatment is of growing
importance.  The key to long-term operating success of such filters
is proper bed design and adequate bed cleaning during backwashing.
A number of questions related to adequate backwashing of granular
filters were investigated and the study results lead to the follow-
ing conclusions:

     Cleaning granular filters by water backwash alone to fluidize
     the filter bed is inherently a weak cleaning method because
     particle collisions do not occur in a fluidized bed and thus
     abrasion between the filter grains is negligible.

     Due to the inherent weakness of water backwashing cited above,
     auxiliary means of improving filter bed cleaning are essential
     for wastewater filters.  Three auxiliary methods were compared
     in a wastewater pilot filtration study.  The most effective
     backwash was provided by air scour and water backwash simul-
     taneously at subfluidization velocities.  The other two aux-
     iliary methods, surface and subsurface wash and air scour
     prior to water fluidization wash were about comparable in
     effectiveness.

     The performance of coarse sand, dual-, and triple-media fil-
     ters was compared, and the backwashing routines appropriate
     for each media are discussed.  A number of investigations con-
     cerning the design and backwashing of dual-media filters are
     also included.

This report was submitted in fulfillment of Grant No. R802140 by Iowa
State University under the partial sponsorship of the U.S. Environ-
mental Protection Agency.  This report covers a period from September 1,
1971 to May 31, 1976, and work was completed as of December 31, 1975.
                                  iv

-------
                          TABLE OF CONTENTS

                                                                   Page

ABSTRACT                                                           iv

LIST OF TABLES                                                     vii

LIST OF FIGURES                                                    xi

NOTATIONS, ABBREVIATIONS AND CONVERSION'FACTORS                    xviii

ACKNOWLEDGMENTS                                                    xxiv

I.    INTRODUCTION                                                  1

II.   CONCLUSIONS                                                   4

      Conclusions Regarding Backwash Effectiveness                  4
      Conclusions Comparing Filter Performance                      5
      Conclusions Regarding Expansion, Intermixing,                 5
        and Dual Media
      Conclusions Regarding Wastewater Filter Design                7

III.  RECOMMENDATIONS                                               10

IV.   BACKWASHING-POTABLE WATER EXPERIENCE                          11

V.    BACKWASHING WITH FLUIDIZATION AND EXPANSION                   16
      Some Fluidization Fundamentals                                16
      Predominance of Hydrodynamic Forces in Cleaning               23
        by Water Fluidization Alone

VI.   OPTIMUM CLEANING BY WATER BACKWASH ALONE                      26
      Particulate Fluidization and Optimum Turbulence -             26
        Evidence from the Literature
      A New Theory of Optimum Backwashing by Water Fluidization     31
        Only
      Experimental Support for the Optimum Theory                   39

VII.  WASTEWATER FILTRATION AND BACKWASHING -                       68
        LITERATURE REVIEW
      Types of Wastewater Filters and Cleaning Techniques           6.8
      Case Histories                                                76

VIII. EXPERIMENTAL COMPARISON OF BACKWASH METHODS IN                89
        WASTEWATER FILTRATION
      Pilot-Plant Equipment                                         g
-------
      Chronology of Experiments                                       91
      Equipment Details                                               91
      Analysis of Samples                                            102
      Operation and Results  - Phase  I Dual-Media                     106
        Filtration of Alum-Treated Secondary Effluent
      Operation and Results  - Phase  II Dual-Media                    126
        Filtration of Secondary Effluent
      Operation and Results  - Phase  III, IV, and V Single-,          157
        Dual-, and Triple-Media Filtration of Secondary
        Effluent
      Operation and Results  - Phase  VI Coarse Sand Filtration        194
        of Secondary  Effluent

 IX.   EXPANSION AND INTERMIXING OF MULTI-MEDIA FILTERS               204
      Introduction                                                  204
      Dual-Media and  Multi-Media Filtration Literature               205
      Backwashing of  Granular Filters                                207
      Bed Expansion Correlations from Fluidization Literature        211
      Prediction of Settling Velocities                              224
      Existing Models for  Predicting the Expansion of Fluidized Beds 230

 X.    EXPANSION AND INTERMIXING EXPERIMENTAL INVESTIGATION           233
      Experimental Apparatus                                        233
      Experimental Procedures                                        239
      Illustrative Calculations                                      242
      Results  and Analysis - Summary                                247
      Results  - Media Characteristics                                247
      Fixed Bed Hydraulic  Profiles in Dual-Media                     256
      Filters - Coal and  Sand
      Expansion - Flow Rate  Observations                             271
      Intermixing Observations

 IX.   EFFECT OF MEDIA INTERMIXING ON DUAL-MEDIA FILTRATION           315
      Introduction                  .                                315
      Objectives and  Scope of this Study                             315
      Experimental Investigation                                     3^
      Results                                                        323
      Discussion                                                    336
      Conclusions                                                    340

XII.  ABRASIVE LOSS OF COAL DURING AIR SCOUR                         341
     Experimental Procedure                                         341
      Results                                                        342
      Conclusions                                                    343

XIII. REFERENCES                                                    345

XIV.  APPENDIX - Sieve Analyses of Uniform Media                     355
                                   vi

-------
                            LIST OF TABLES

                                                                     Page

Table 1.   Analysis of University tap water.                          43

Table 2.   Experimental design for series 2.                          47

Table 3.   Backwash procedures for graded sand.                       48

Table 4.   Effluent quality index and porosity for series 2.          54

Table 5a.  Manual backwashing procedure used on full-scale rapid      83
           sand filters at Luton [86].

Table 5b.  Backwashing procedure in full-scale rapid sand filters     83
           at Luton in November 1975.

Table 6a.  Automatic backwashing sequence used on pilot scale         84
           Immedium upflow filter at Luton [86].

Table 6b.  Backwash procedure for full-scale upflow filters at        84
           Luton in November 1975.

Table 7.   Summary of experimental phases for wastewater filtration   92
           backwashing study.

Table 8.   Filter media details for wastewater filtration pilot       99
           studies.

Table 9.   Initial filter head losses during various portions of     113
           the study.

Table 10.  Backwash rates required to achieve 38 to 40% bed          120
           expansion.

Table 11.  Results of analyses during alum treatment (Phase I)       121
           for samples from May 17 to July 11, 1973, when both
           filters were washed by water fluidization only.  (All
           results from composite samples except solids contact
           influent.)

Table 12.  Results of analyses during alum treatment series          122
           (Phase I) for samples from July 11 to August 20, 1973,
           when air scour was being used on the south filter.
           (All results from composite samples except solids
           contact influent.)

Table 13.  Suspended solids released from filters in special         125
           backwashes after run 78.

Table 14.  Summary of head loss development during observation       150
           runs of Phase II, during direct filtration of
           secondary effluent.
                                   vii

-------
Table  15.  Results of  analyses during direct filtration of            151
           secondary effluent  (Phase II) from August 30 to
           October 30, 1973.   (All composite samples except
           as noted.)

Table  16.  Data  summary of clean-up operation at the end of           153
           Phase II.

Table  17.  Solids capture per unit head loss results for direct       183
           filtration  of trickling filter effluent, 1974.

Table  18.  Summary of  analytical test results for Phase III.          185

Table  19.  Summary of  analytical test results for Phase IV.           186

Table  20.  Summary of  analytical test results for Phase V.            187

Table  21.  Data  summary for clean-up operation at the end of          189
           the operation period in 1974.

Table  22.  Action of simultaneous air and water backwash on           197
           coarse sand at subfluidization velocities.

Table  23.  Results of  analyses during direct filtration of            199
           secondary effluent (Phase VI) from June 24 through
           August 2, 1975, using coarse sand filters of
           different depths.

Table  24.  Mean  total head loss during filtration of secondary        200
           effluent on coarse sand filters during Phase VI.

Table  25.  Average initial head loss for three coarse sand filters    200
           before and  after run 35 in Phase VI when increase of
           backwash rate was adopted.

Table  26.  Height that sand is thrown by simultaneous air and         202
           water backwash.

Table  27.  Comparison of n slopes of sea sands (using Jottrand's      217
           analysis [69]  and Richardson and Zaki's equations).

Table  28.  Comparison of n slope of crushed coal.                     218

Table  29.  Size and source of raw graded silica sands and             238
           coal studies.

Table  30.  Sieve analysis of garnet sand media (-14+16).              243
                                  viii

-------
                                                                     Page
Table 31.  Upflow and/or downflow experimental runs with the         248
           various single media in the 6-in. column.

Table 32.  Upflow and downflow experimental runs with dual-          249
           media filters in the 6-in. column.

Table 33.  Upflow experimental runs with various uniform media       250
           and uniform dual media in 2-in. column, 25 °C.

Table 34.  Average diameter of uniform media by two methods, mean    254
           of adjacent sieve sizes and mean equivalent spherical
           diameter by the count and weigh method [Eq. (40)].

Table 35.  Summary of the average diameters of the media - d         255
           (by several methods).

Table 36.  Densities of media - p                                    256

Table 37.  Fixed-bed porosities of the three media determined        257
           by the two techniques (e ) and two investigators.

Table 38.  Settling velocities of uniform garnet and sand media.     258

Table 39.  Expansion - flow rate data of run 1, Series A-13          272
           (-14+16 garnet and sand media).

Table 40.  Summary of minimum fluidization velocities of garnet      285
           sand media - V c.
                         mf
Table 41.  Results of the log V vs log e relationship for            289
           garnet sand media.

Table 42.  Values of Reynold's numbers and Galileo's number for      291
           the garnet sand media.

Table 43.  V., n, e f to be used in author's expansion models.       292

Table 44.  Results of log V vs log e regression analyses for         293
           uniform sized silica sands and coals.

Table 45.  Predicted values of m, VA with errors of prediction       295
           and maximum error of prediction of expanded bed
           depth for uniform sands.

Table 46.  Predicted values of n, V^ with errors of prediction       296
           and maximum error of prediction of expanded bed
           height for uniform coals.
                                    ix

-------
Table 47.  Prediction of expanded bed depths for graded
           sand using models developed for uniform sands
           and average diameter based on the inverse
           definition [Eq. (38)].

Table 48.  Bulk density difference, lb/ cu ft (garnet-silica
           sand).

Table 49.  Identification of experimental series.

Table 50.  Influent and effluent suspended solids data, average
           and range, for wastewater series.

Table 51.  Average values of background (bg) turbidity for the
           treated wastewaters.

Table 52.  Summary of results comparing filter performance for
           sharp and mixed interface media.

Table 53.  Changes in coal bed over a two-week period of air-
           scour exposure equivalent to about 20 years of
           normal service.
Page

298
307


324

335


336


336


343

-------
                           LIST OF FIGURES

                                                                     Page

Fig. 1.   Characteristics of fluidized beds.                          17

Fig. 2.   Fundamental behavior patterns.                              19

Fig. 3.   Shear forces on an elemental volume of fluid.               33

Fig. 4.   Schematic layout of experimental apparatus.                 40

Fig. 5.   Head loss curves for run 3B, top 6 in. of filter media.      51

Fig. 6.   Variation of the ratio of effluent to influent iron         52
          with time.

Fig. 7.   Cumulative differential effluent iron vs time, run 8.       53

Fig. 8.   Cumulative effluent quality index vs porosity, series 2,    55
          12-in. depth.

Fig. 9.   Cumulative effluent quality index vs porosity, series 2,    56
          all depths.

Fig. 10.  Cumulative effluent quality index vs porosity, series 1,    56
        * 12-in. depth.

Fig. 11.  Cumulative effluent quality index vs expansion, series 3,   57
          18-in. depth.

Fig. 12.  Cumulative effluent quality index vs expansion, series 3,   58
          all depths.

Fig. 13.  Backwash water quality vs washwater volume, series 1.       59

Fig. 14.  Backwash water quality vs washwater volume, series 1.       60

Fig. 15.  Backwash water quality vs washwater volume, series 1.       61

Fig. 16.  Terminal backwash water quality vs porosity, series 1.      62

Fig. 17.  Backwash water volume vs porosity, series 1.                64

Fig. 18.  Iron removable by physical abrasion test vs expansion,      65
          runs 20 and 21.

Fig. 19.  Pilot-scale Immedium filter used at West Hertfordshire,      71
          England [142].

Fig. 20.  Immedium filter arrangements for full-scale installations,  72
          open and pressure [13].


                                    xi

-------
 Fig.  21.   Environmental Elements Corp.  (Koppers) full-scale           75
           automatic  backwash  filter  (from manufacturer's brochure).

 Fig.  22.   Pilot-scale  "Simater" radial-flow, moving-bed sand          77
           filter by  Simonacco Ltd. of Carlisle, England [68].

 Fig.  23.   Schematic  representation of pilot-scale filter plant        90
           used  in  experimental investigation.

 Fig.  24.   Pilot plant  solids contact unit.                            94

 Fig.  25.   Details  of filter boxes.                                    96

 Fig.  26.   Abrasion test when filtering  secondary effluent treated    112
           with  alum  for phosphorus reduction in Phase I.

 Fig.  27.   Head  loss  vs time at various media depths, run 27.         115

 Fig.  28.   Head  loss  vs time at various media depths, run 42.         116

 Fig.  29.   Head  loss  vs time at various media depths, run 59.         117

 Fig.  30.   Head  loss  vs time at various media depths, run 63.         118

 Fig.  31.   Head  loss  vs time at various media depths, run 71.         119

 Fig.  32.   Suspended  solids concentration of backwash water vs        124
           quantity of backwash water used, run 27, second back-
           wash  of  the south filter immediately following the
           first application of air scour.

 Fig.  33.   Standardized abrasion test results (Phase II) during       136
           direct filtration of secondary effluent.

 Fig.  34.   Initial head loss data for north, south, and west          138
           filters  for entire Phase II study during direct
           filtration of secondary effluent.

 Fig.  35.   Frequency  plot of initial head loss data, phase II.        139

 Fig.  36.   Chronological head loss development at various media       141
           depths, run 2.

 Fig.  37.   Chronological head loss development at various media       142
          depths, run 14.

Fig.  38.  Chronological head loss development at various media       143
          depths, run 22.

Fig.  39.  Chronological head loss development at various media       144
          depths, run 29.
                                   xii

-------
Fig. 40.  Chronological head loss development at various  media       145
          depths, run 36.

Fig. 41.  Chronological head loss development at various  media       146
          depths, run 43.

Fig. 42.  Chronological head loss development at various  media       147
          depths, run 54.

Fig. 43.  Chronological head loss development at various  media       148
          depths, run 64.

Fig. 44.  Results of special backwash, day no.  242.

Fig. 45.  Initial head loss in bottom 16 in.  of coarse media         168
          filter, observation runs only.

Fig. 46.  Standard abrasion test results for  Phase III.              17°

Fig. 47.  Standard abrasion test results for  Phase IV and V.         171

Fig. 48.  Initial head loss data for each filter for entire study.    173

Fig. 49.  Chronological head loss development at various  media       176
          depths, day no. 155, Phase III.

Fig. 50.  Chronological head loss development at various  media       177
          depths, day no. 183, Phase III.

Fig. 51.  Chronological head loss development at various  media       178
          depths, day no. 239, Phase IV.

Fig. 52.  Chronological head loss development at various  media       179
          depths, day no. 240, Phase V.

Fig. 53.  Chronological head loss development at various  media       180
          depths, day no. 267, Phase V.

Fig. 54.  Chronological head loss development at various  media       181
          depths, day no. 269, Phase V.

Fig. 55.  Chronological head loss development at various  media       182
          depths, day no. 281, Phase V.

Fig. 56.  Relationship between superficial velocity - porosity.       212

Fig. 57.  Relationship between n slope and Reynolds number Re .       214
                                                             o

Fig. 58.  Schematic layout of 6-in. fluidization column.              234

Fig. 59.  Schematic layout of 2-in. fluidization column.              236
                                  xiii

-------
Fig. 60.  Sieve analysis of graded sand media.                       251

Fig. 61.  Sieve analysis of graded coal media.                       251

Fig. 62.  Sieve analysis of garnet sand media.                       252

Fig. 63.  Fixed-bed head loss for graded sand A at 9 gpm/sq ft       259
          for various temperatures.

Fig. 64.  Fixed-bed head loss for graded sand A at 26.5 °C for       260
          various flow rates.

Fig. 65.  Head loss for individual graded media in 1-1/2 in. unit    262
          filter sections.

Fig. 66.  Head loss for individual graded media in 1-1/2 in. unit    263
          filter sections.

Fig. 67.  Fixed-bed head loss of dual media AA at various tempera-   264
          tures and flow rates.

Fig. 68.  Head loss per 1-1/2-in. unit depth in dual media AA and    265
          head loss for the two-component media if unmixed.

Fig. 69.  Head loss per 1-1/2-in. unit depth in dual media AC and    266
          head loss for the two-component media if unmixed.

Fig. 70.  Head loss per 1-1/2-in. unit depth in dual media AD and    267
          head loss for the two-component media if unmixed.

Fig. 71.  Head loss per 1-1/2-in. unit depth in dual media A2E       268
          and head loss for the two-component media if unmixed.

Fig. 72.  Head loss per 1-1/2-in. unit depth in dual media AF        269
          and head loss for the two-component media if unmixed.

Fig. 73.  Pressure loss - flow rate diagram for garnet sand media    273
          (-14+16).

Fig. 74.  Pressure loss - flow rate diagram for garnet sand media    274
          (M-60-80).

Fig. 75.  Expansion - flow rate characteristics (garnet sand, run    275
          1).

Fig. 76.  Expansion - flow rate characteristics (garnet sand, run    276
          2).

Fig. 77.  Expansion - flow rate characteristics (garnet sand, run    277
          3).
                                   xiv

-------
Fig. 78.  Expansion - flow rate characteristics (garnet sand, run    278
          4).

Fig. 79.  Expansion - flow rate characteristics (garnet sand, run    279
          5).

Fig. 80.  Expansion - flow rate characteristics (garnet sand, run    280
          6).

Fig. 81.  Expansion - flow rate characteristics (garnet sand, run    281
          7).

Fig. 82.  Expansion - flow rate characteristics (garnet sand, run    282
          8).

Fig. 83.  Expansion - flow rate characteristics (garnet sand, run    283
          9).

Fig. 84.  Expansion - flow rate characteristics (garnet sand, run    284
          10).

Fig. 85.  Log plot of V vs e for garnet sand media (-14+16) (run     286
          1, Series A-13).

Fig. 86.  Log plot of n slope vs Reynold's number - Re^ (for         287
          garnet sand media, runs 1 through 10, Series A-13
          through A-17).

Fig. 87.  Log plot of Reynold's number, Re^, vs Galileo number, Ga   290
          (for garnet sand media, runs 1 through 10, Series A-13
          through A-17).

Fig. 88.  Minimum fluidization velocity, Vmf, to achieve 10% bed     300
          expansion at 25 °C.

Fig. 89.  Effect of temperature on Vmf for sand and coal and on      300
          absolute viscosity of water.

Fig. 90.  Bulk density vs flow rate for garnet sand and silica       302
          sands.

Fig. 91.  Intermixing of -50+60 garnet sand and -20+25 silica        303
          sand.

Fig. 92.  Intermixing of -50+60 garnet sand and -30+35 silica        303
          sand.

Fig. 93.  Intermixing of -50+60 garnet sand and -35+40 silica        304
          sand.

Fig. 94.  Intermixing of -50+60 garnet sand and -40+45 silica        304
          sand.
                                   xv

-------
 Fig. 95.  Intermixing of silica sand and coal according to the      309
           model of Camp et al. [26].

 Fig. 96.  Bulk density vs flow rate for coal and silica sands       310
           (data points not shown on all curves for drafting
           convenience).

Fig.  97.  Expansion vs flow rate of dual media AA and the two-      312
           component media at 22 °C.

Fig.  98.  Expansion vs flow rate of dual media A2C2 and the         314
           two-component media at 22 °C.

Fig.  99.  Schematic diagram of apparatus.                           317

Fig. 100.  Head loss and filtrate quality vs volume of filtrate,     325
           Series IS, run 2, sharp interface, filtration of iron
           with CQ = 8.68 to 9.64 mg/1 Fe, Avg 9.07 mg/1.

Fig. 101.  Head loss and filtrate quality vs volume of filtrate,     326
           Series IM, run 11, mixed interface, filtration of iron
           with CQ = 9.2 to 9.7 mg/1 Fe, Avg 9.42 mg/1.

Fig. 102.  Head loss and filtrate quality vs volume of filtrate,     327
           Series II S, run 1, sharp interface, filtration of
           activated sludge effluent with C  = 4.5 to 8.5 FTU,
           Avg 6.16 FTU.                   °

Fig. 103.  Head loss and filtrate quality vs volume of filtrate,     328
           Series II M, run 1, mixed interface filtration of
           activated sludge effluent with C  = 2.6 to 8.6 FTU,
           Avg 4.15 FTU.                   °

Fig. 104.  Head loss and filtrate quality vs volume of filtrate,     329
           Series III  S, run 3,  sharp  interface, filtration of
           alum coagulated  trickling filter  effluent C   - 3.8  to
           10 FTU,  Avg 5.21 FTU.                      °

Fig. 105.  Head loss and  filtrate quality vs volume of filtrate,     330
           Series  III  M,  run 5,  mixed  interface, filtration of
           alum coagulated  trickling filter  effluent with CQ  =  1.7
           to 6.2 FTU, Avg  2.95  FTU.

Fig. 106.  Head loss and  filtrate quality vs volume of  filtrate,     331
           Series  IV S, run 3, sharp interface,  filtration of
           trickling filter effluent with CQ = 4.7 to  8.5 FTU,
           Avg  6.29 FTU.

Fig. 107.  Head loss and filtrate quality vs volume of filtrate,     332
           Series  IV M, run 3, mixed interface,  filtration of
            trickling  filter effluent with CQ = 12  to  15  FTU,
           Avg 12.6 FTU.


                                   xvi

-------
Fig. 108.  Headless and filtrate quality vs volume of filtrate,      333
           Series V S, run 4, sharp interface, filtration of
           limesoda ash softening precipitate with C  = 5.8 to
           6.5 FTU, Avg 6.14 FTU.

Fig. 109.  Head loss and filtrate quality vs volume of filtrate,     334
           Series V M, run 3, mixed interface, filtration of lime-
           soda ash softening precipitate with C  = 6.8 to 8.3
           FTU, Avg 7.5 FTU.                    °

Fig. 110.  Sieve analysis of coal before and after 14 days of        344
           continuous air scour.
                                  xvii

-------
          NOTATION, ABBREVIATIONS AND CONVERSION FACTORS





B         = Camp's backwash ing number



C         - volume



C         • influent  concentration
 o


C         = effluent  concentration



C/C       = effluent  concentration/ influent concentration
   o


Cn        = drag coefficient



d         = particle  diameter                                  L



D         = diameter  of column or bed                          L



d.        = diameter  of particle in  'i'th layer                L



d         = equivalent diameter of spherical particle          L
d607 finer= 60^° finer of a particle from a probability plot    L



d.        = l/£(w./d.) = average diameter of particle by

            inverse definition                                 L



                   arithmetic mean diameter                    L
                                                               _2
F.        = buoyant force                                   ML T


                                                               _2
F£        = impelling force                                 ML T


                                                               -2
g         * acceleration due to gravity                     LT


    dV1                                                      -1
G = -: —   = mean velocity or shear gradient in pores        T
    CLZ

             3            2
Ga        = d p(p  - p)g/fa  Galileo number
                 s


Gf        = superficial fluid mass velocity, Ib(mass)/         „  -

            hr sq ft                                        ML T



G ,       = superficial fluid mass velocity at minimum

 m          fluidization, Ib (mass)/hr sq ft                ML T



h- or HL   = head loss in flow through granular bed             L



i         = subscript denoting the 'i'th layer of the bed
                               xviii

-------
K         = f(Vs, \|f, d/Dt) = constant for a particular      LT~
            fluidized system

A         = height of bed                                      L

SL         = height of static bed                               L
 o

m         = index of fluid regime

mf        = subscript denoting condition at minimum
            fluidization

n         = slope of log V vs log e plot

N         = number of particles
                                                              -1 -2
p         = pressure intensity                              ML  T

                                                              2 -3
P         = power dissipated                                ML T

r         = d /d  ratio of the particle diameters of
             x  y
            x and y components

R         = resistance force per unit projected area of     ML  T
            the particle

Re        = pV d/M, = Reynold's number based on the
            superficial velocity of the fluid above the bed

Re.       = Reynold's number based on the velocity V^
            intercept at porosity equals one of the log V vs  log
            e plot

Re        =  pVgd/n = Reynold's number based on unhindered
  0         settling velocity of particle

                                                              -1 -2
S         = mean shear stress                               ML  T

t         = time                                               T
                                                                3
v         = volume of particle                                 L

V         = superficial velocity of the fluid               LT~
            above the bed

V.        • velocity intercept at a porosity ratio of
            one of the log V vs e plot, equal to  the
            settling velocity of a discrete spherical          ,
            particle (Vs) when d/D is negligible            LT
                                xix

-------
V f       = minimum fluidization or critical fluid
            velocity expressed as a superficial               _^
            velocity                                        LT

V         = unhindered settling velocity of a                 _j_
            discrete particle                               LT

V'        = V/e = average fluid velocity within                ,
            pores of filter                                 LT

W         - total weight of particles

W.        = weight fraction of  'i'th layer

          = Cartesian coordinates                              L

                          Greek Symbols
                        I In
a         = [MgK(pg-p>]  '   = constant for a particular       _j_ _2
            fluidized system in optimum backwashing theory  ML  T
                                                              -2 -2
Y         = specific weight of  fluid                        ML  T

\        -  (Ps  -  P)/
               x         y                                    -2-2
Y0        =  specific weight of  particle                     ML  T
  a

 6        = x/d  =  dimensionless spacing of  particles
             in fluidized  state

A        =  prefix signifying  increment

 e        =  porosity  ratio

 ee       = porosity  ratio  of  expanded bed

 e £       = porosity  ratio  at  minimum fluidization
             velocity

 e         = porosity  ratio  of  fixed bed
                                                             ,,-lnT1
 |a         = viscosity of fluid in centipoise                ML  T
                                                              2  -1
 V         = H/P  » kinematic viscosity                       L T
                                                               -3
 p         = fluid density                                   ML
                                                               -3
 p,         = (1 - e)p   + pe =  bulk density of mixture       MT
                                                               .3
 p         = particle density                                ML
  s
                                 xx

-------
            sphericity ratio of the surface area of
            an equivalent volume sphere to the actual
            surface area of the particle
                           Abbreviat ions

American Society for Testing Materials                      ASTM

biochemical oxygen demand                                   BOD

centimeter                                                  cm

chemical oxygen demand                                      COD

cubic centimeter                                            cc

cubic feet                                                  cu ft

cubic feet per minute                                       cfm

cubic feet per second                                       cfs

degree(s) Celsius                                            C

effective size                                              ES

degree(s) fahrenheit                                         F

feet                                                        ft

feet per second                                             fps

Formazin turbidity units                                    FTU

gallon(s)                                                   gal.

gallon(s) per minute                                        gpm

gram(s)                                                     g

horsepower                                                  HP

hour(s)                                                     hr

inch(es)                                                    in.

inches per minute                                           ipm
                               xxi

-------
inside diameter                                              ID




Jackson turbidity units                                      JTU




mercury                                                      Hg




micrometer(s)                                                pro




microliter(s)                                                Ml




million gallon(s)                                            MG




million gallons  per  day                                      MGD




milligram(s)  per litre                                      mg/1




milliliter(s)                                               ml




millimeter(s)                                               nnn




mimute(s)                                                   min




 nanometer                                                   nm




 number                                                      no •




 outside diameter                                            OD




 parts per million                                           PP™




 parts per trillion                                          PPfc




 percent                                                     %




 pound(s)                                                     1°




 pounds per square inch,  absolute                             psia




 pounds per square inch,  gage                                 psig




 revolutions  per minute                                       rPm




 root mean  square                                            rms




  second(s)                                                    sec




  square  centimeter                                            S<1 cm




  square  feet                                                  sq ft




  suspended solids                                            ss
                                xxii

-------
 trickling filter
 uniformity coefficient
 versus
 volume
 weight
                      Conversion Factors
                                    TF
                                    UC
                                    vs
                                    vol
                                    wt
The following  factors convert  the units used  in  this  report  to  the
more common SI metric unit  in  popular usage in water  engineering
practice.
   English Unit     x

cfm(cu ft/min)
cfm/sq ft

ft
fps (ft/s)
gal
gal/sq ft
gpm (gal/min)
  it    it
gpm/sq ft

g/sq ft (gram/sq ft)
hp
in.
in./min
Ib
Ib/cu ft
Ib/hr ft
Ib/hr sq ft
mgd (million gal/day)
psi
sq ft
Multiplier       -      Metric Unit (SI)

  1.68                      m3/h
  18.288                    m3/m2h
           (i.e., m/h superficial velocity)
  0.3048                    m
  0.3048                    m/s
  .003785                   m3
  .0407                     m3/m2
  0.2272                    m3/h
  0.0631                    1/s
  2.442                     m3/m2 h
           (i.e., m/h superficial velocity)
  10.750                    g/m2
  0.7457                    kw
  0.0254                    m
  1.524                     m/h
  0.454                     kg
  16                        kg/m3
  1.489                     kg/h m
  4.88                      kg/h m2
  3785                      m3/d
  6.9                       kN/m2
  0.0929                    m2
                                xxiii

-------
                        ACKNOWLEDGMENTS
The research reported herein was supported by the Engineering Research
Institute at Iowa State University, Ames, Iowa, in part through funds
made available by research grant No. R802140 (formerly 17030 DKG)
from the Office of Research and Monitoring of the U.S. Environmental
Protection Agency.

The report incorporates the work of several graduate student theses
in sanitary engineering at Iowa State.  The students were A.
Amirtharajah, R. R. Boss, W. J. Carvalho, J. C. Lorence, A. M. Malik,
G. A. Rice, G. D. Sejkora, E.  W.  Stangl,  and C. F. Woods.   Their
contribution is gratefully acknowledged.  The report was compiled by
the principal investigator, John L. Cleasby.  The assistance of
Oliver Hao in the statistical analysis of the data is also
acknowledged.
                                xxiv

-------
                           I.   INTRODUCTION
 The use of deep granular filters in wastewater treatment  is  of
 growing importance.   Filters are an essential unit  operation in
 many tertiary wastewater treatment flow schemes and in  all physical-
 chemical wastewater  treatment flow schemes.

 The deep granular filter used in wastewater  treatment is  subjected
 to more severe operating conditions than it  faced in potable water
 treatment.  The wastewater filter receives heavier  and more  varia-
 ble influent suspended-solids loads,  and the solids tend  to  stick
 more tenaciously to  the filter media.

 The key to long-term operating success  of deep granular filters  is
 proper bed design and adequate bed cleaning  during  backwashing.
 Due to the heavier burden received by such filters, a coarser media
 at the entering surface is essential to achieve reasonable filter
 run length.   Dual- and triple- (multl-) media are commonly being
 used to achieve the  coarser surface media in the United States.
 Coarser coal sizes are advocated to encourage better penetration of
 suspended solids (and thus longer filter runs),  preventing the
 ready transfer of prior experience from the  water filtration field
 to wastewater filtration.   Other approaches, such as deep beds of
 coarse sand backwashed without bed expansion and shallow  beds of
 fine sand backwashed automatically at frequent intervals  are also
 being used.

 The research conducted on this grant was designed to answer  a num-
 ber of important questions related to the design and operation of
 wastewater filters.   In general,  the original questions proposed
 for study were how to select the appropriate media  sizes  in  dual-
 and triple-media filters,  how to predict the minimum and  optimum
 backwash flow rate for the media selected, and how  to demonstrate
 the value of air scour or  surface wash  as auxiliaries to  water
 backwashing.

 Recent developments  have added even greater  importance to the proj-
 ect,  as discussed in the following paragraphs.

The importance  of wastewater  filters is  emphasized by the require-
ment of  secondary  treatment as the minimum acceptable treatment and
by the recent EPA definition  of secondary treatment which requires
an  effluent quality of  30 mg/1 BODs and 30 mg/1 suspended solids,
both for  30 consecutive  day averages.  Many existing municipal
plants cannot meet the 30 BOD, 30 suspended solids goal, especially
trickling filter plants  during winter seasons.  Tertiary filtration
of such  effluents appears  to  be one of the most attractive alterna-
tives for upgrading such plants to meet  the new standards.  This
attractiveness  exists because  the technology is well known from

-------
 long use  in  potable water treatment,  and because the operating
 costs  and energy  requirements are  low.

 Tertiary  wastewater filtration is  also attractive for plants which
must meet more stringent effluent  standards.  In some locations,
 effluent  BOD5 and suspended solids limits of 5 mg/1 each have been
 established.  Direct filtration of a  good secondary effluent will
 come close to meeting this goal.   Chemical coagulation prior to
 filtration will be needed in most cases.

 Thus,  it  appears  that a growing number of tertiary filtration
 plants will  be installed in the next  few years.  It is vitally im-
 portant that tertiary filters be designed and built so they do not
 fail to meet the desired standards over their service life.  A
vital  element of proper design is  the provision of adequate back-
washing facilities.

The growing  interest in wastewater filtration is evidenced by a.
 growing number of equipment companies marketing such equipment.
 Some of these companies know very  little about filtration, espe-
cially wastewater filtration.  The interest is also evidenced by
 requests  from EPA regional offices for Technology Transfer Seminars
 on upgrading secondary plants by tertiary filtration.

 During the third year of the current project, the crucial impor-
 tance of  the backwashing provisions for tertiary wastewater filters
was demonstrated and is reported herein.  Water fluidization alone
 as a means of backwashing was found to be totally inadequate to
maintain  the filter media in acceptable condition.   Both the use of
an air-scour auxiliary before the water backwash and the use of a
surface wash auxiliary before and during the water backwash made
substantial  improvement in the cleaning effectiveness.   However,
some evidence of inadequate backwash remained, even with these
auxiliary cleaning schemes.

                           Specific Aims

The specific aims of the project from the original research propos-
al are repeated here for reference, with modifications  as adopted
in the continuation proposals.

A-l  To attempt to extend the successful prediction model of
     Amirtharajah for sandexpansion to anthracite coal  and garnet
     sand filter media.   If  unsuccessful,  attempt to develop a
     satisfactory model.

B-l  To test the validity of the equal bulk density approach to
     prediction of degree of intermixing of two filter media of
     different size and specific gravity.

-------
C-l  To evaluate the hydraulic behavior of dual-  and multi-media
     beds with different degrees of interfacial intermixing  and
     develop a tentative bed design approach.

C-2  To compare the filtering efficiency of a partially intermixed
     filter bed with the same media without intermixing.   Also, to
     compare the ease of backwashing of the two types  of filter bed.

C-3  To compare the effectiveness of backwash at  various rates pro-
     viding different degrees of expansion, and to attempt to veri-
     fy or refute the theoretical optimum expected at  porosities
     of 0.65 to 0.70.  Backwash effectiveness will be  measured
     first by the bed cleanliness resulting from the wash and later
     by the quality of the water in the next filter run.

C-4  To compare the effectiveness of a programmed backwash covering
     a range of wash rates with a constant rate backwash, and to
     recommend an optimum water backwashing procedure.  This objec-
     tive was dropped in the second year continuation  proposal.

C-5  To compare the effectiveness of various air-water backwashing
     schemes with the optimum water wash procedure alone.  This ob-
     jective was expanded to include surface wash auxiliary in the
     second year of the study.

C-6  To reaffirm or revise the tentative design of dual- and multi-
     media filters suggested in (C-l).

Additional objectives were added in the third year continuation
proposal as follows:

C-7  To determine if coarse, single-media, sand filters can be
     cleaned successfully by using air and water simultaneously,
     with the water backwash rates well below minimum fluidization
     velocity for the media.

C-8  To evaluate the acceptability of filter designs which permit
     backwashing with various qualities of backwash water, including
     the feed water (secondary effluent).

The original objectives were heavily oriented to dual- and triple-
media filters because that is the common approach in the United
States.  During the conduct of the project over a four-year time
period, it became evident that other types of granular filters may
have merit, and some new objectives were added dealing with the
deep-bed coarse sand filter.  No work has been included in this
study on proprietary filters such as the ABW filter, the Hydroclear
filter, the Immedium upflow filter, etc.  The absence  of such work
should not be taken as either a favorable or unfavorable view
toward such filters.  Rather, their omission from the  study only
reflects that limited objectives had to be selected to fit within
the scope of the project.

-------
                         II.  CONCLUSIONS
A four-year study of wastewater filtration and filter backwashing
is reported herein.  Various granular media filters were studied
including those using single-, dual-, and triple-media.  Various
methods of backwashing were compared including:  (1) water fluidi-
zation only, (2) air scour followed by water fluidization, (3) sur-
face wash and subsurface wash before and during water fluidization
backwash, and (4) simultaneous air scour and subfluidization water
backwash.

           Conclusions Regarding Backwash Effectiveness

 1.  The cleaning of granular media filters by water backwash alone
     to fluidize the filter bed is an inherently weak cleaning
     method because particle collisions do not occur in a fluidized
     bed and thus abrasion between the filter grains is negligible.

 2.  The cleaning which results in a water fluidized bed is due to
     the hydrodynamic shear at the water-media grain interfaces.  A
     simple mathematical model was developed to calculate the poros-
     ity of maximum hydrodynamic shear in a fluidized bed.  Maximum
     hydrodynamic shear in a fluidized bed occurs at a porosity of
     0.68 to 0.71 for sand sizes normally used in filtration.  Opti-
     mum cleaning of the filter media at this porosity was demon-
     strated experimentally.

 3.  When backwashing by water fluidization alone, a slight economy
     in total washwater used is achieved by expanding the bed to
     the optimum porosity outlined in conclusion 2 above.  Lower
     wash rates (anywhere above the rate for minimum fluidization)
     will result in nearly the same terminal washwater turbidity,
     but proportionately longer backwash times will be required.
     Therefore no economy of water use is achieved by use of low
     backwash rates.

 4.  The weakness of water fluidization backwash alone was clearly
     demonstrated during wastewater filtration studies where a
     dual-media filter which was washed by water fluidization alone
     developed serious dirty filter problems such as floating mud
     balls, agglomerates at the walls, and media surface cracks.
     These problems were observed when filtering either secondary
     effluent or secondary effluent which had been treated with
     alum for phosphorous reduction.

 5.  The heavy mud ball and agglomerate accumulations caused higher
     initial head losses and shorter filtration cycles.  They may
     also cause  poorer filtrate quality in some cases, although
     such detriment was not demonstrated in this study.

-------
  6.   Simultaneous air  scour  and  subfluidization backwash of coarse
      sand  filters proved  to  be the most effective method of backwash.
      However,  this method should not be used for finer filter media
      such  as  the coals and sands of  the typical sizes used in dual-
      and triple-media  filters because  loss of media will occur during
      backwash  overflow.   The choice  of the simultaneous air and water
      flow  rates must be appropriate  for the size of sand being used,
      and should result in some circulation of the sand for effective
      backwashing.

  7.   The other two methods of improving backwashing, namely air
      scour followed by water fluidization backwash, and surface
      (and  subsurface)  wash before and  during water fluidization
      backwash, proved  to  be  comparable methods of backwash which
      can be applied to single-,  dual-  and triple-media filters.
      These two methods did not completely eliminate all dirty fil-
      ter problems, but both  auxiliaries reduced the problems to
      acceptable levels so that filter  performance was not impaired.

             Conclusions  Comparing Filter Performance

  8.   In a  comparison of a coarse sand  filter (2 to 3.6-mm sand,
      46 in. deep)* with a dual-media filter  (1  to 2-mm coal, 15  in.
      deep; 0.5 to 0.8-mm  sand, 9 in. deep) and a triple-media fil-
      ter (same as dual except for the  addition of 3 in. of 0.27 to
      0.54-mm garnet sand)  while  filtering secondary effluent, the
      coarse sand filter produced a filtrate slightly poorer in
      quality,  but produced substantially more filtrate to a common
      terminal  head loss.   The performance of the dual- and triple-
      media filters was comparable.

  9.   In a  comparison of three coarse media sand filters (2.5 to
      3.7-mm sand) of different depths  (24, 47, 60 in.) while fil-
      tering secondary  effluent,  there  was no apparent difference in
      filtrate  quality  or  in  rate of head loss development.

                 Conclusions Regarding Expansion,
                    Intermixing,  and Dual Media

 10.   Methods for prediction  of filter  bed expansion are desirable
      for rational design  of  filters  and filter backwashing provi-
      sions.  Existing  models for prediction of filter bed expansion
      are not adequate  for the three  filter media in prominent use
      today;  coal, silica sand,  and  garnet sand.  New unified
*
 This report contains common English units since part of the data are
 derived from progress reports prepared before the requirement for
 reporting in metric units.  Metric equivalents will be found on page
 xxvi

-------
     empirical models for prediction of bed expansion for all
     three common media are presented.

11.  The expansion prediction models provide acceptable prediction
     accuracy for garnet and silica sand, but the coal model is not
     sufficiently accurate without further refinement.  Some mea-
     sure of sphericity for coal seems essential to improvement of
     the model.  In the expansion models, prediction is made from
     measurable properties of the media and the fluid including
     media density, fixed bed porosity, sieve analysis, fluid den-
     sity and fluid viscosity.  The models permit calculation of
     expanded bed porosity, bulk density, and bed depth as a func-
     tion of backwash flow rate.

12.  Complete fluidization of the filter bed during backwashing is
     common practice in the United States.  Data are presented for
     the minimum fluidization velocity of all three media of vari-
     ous common sizes.  The data will be useful in selecting mini-
     mum backwash rates for specific media and water temperatures.

13.  The prediction of intermixing tendency between different fil-
     ter media in use today is important to rational design of dual
     and triple media filters.  Lack of such capability may lead to
     filters with the media excessively intermixed, or with the
     fine media located on the top of the filter bed where it is
     not wanted.  Two existing models for prediction of intermixing
     between media of different size and specific gravity .were
     tested against experimental data.  Both were found to be lack-
     ing in sensitivity.

14.  The degree of intermixing is a function of the backwash flow
     rate and the manner in which the backwash valve is shut off at
     the end of the filter backwash.

15.  There is an intermixing tendency between garnet sand and silica
     sand, and between silica sand and coal; intermixing in both
     tends to increase at higher backwash rates.  This is because
     the bulk density of the heavier media,  in each case, decreases
     more rapidly than that of the lighter media as flow rate is
     increased.

16.  Under some  circumstances,  the smaller but higher specific
     gravity media can be on the bottom of the bed at lower back-
     wash rates  but move to the top of the bed at higher backwash
     rates.   Recommended guidelines for differences in bulk density
     and ratios  of diameter between garnet and silica sand to pre-
     vent this occurrence are presented.

-------
17.  Rapid shut off of the backwash valve tends to trap the differ-
     ent media in the general position associated with the maximum
     backwash rate.  Very slow closing over several minutes would
     be required for restratification of the bed in new positions
     associated with lower backwash flow rates.

18.  Intermixing at the interface of typical dual-media filters
     comprised of coal and silica sand decreases the permeability
     of the lower coal layers and increases the permeability of  the
     upper sand layers.  Thus, the intermixed zone has pore diam-
     eter, permeability,  and fixed-bed hydraulic gradient values
     during flow which lie between the values for those parameters
     for the individual media.  This tends in the desired direction
     of providing coarse  to fine filtration.

19.  The degree of intermixing at the interface of coal and silica
     sand filters is determined more by the amount and size of the
     finer sands present  than by the coal.

20.  Interfacial intermixing of dual-media filters does not in it-
     self affect filter performance as measured by both head loss
     development and effluent quality.

21.  Intermixing at the interface of dual-media filters is an un-
     avoidable phenomenon which results when United States anthra-
     cite coal (sp gr about 1.7) and sand are used and the sizes
     are selected to achieve coarse to fine filter media in the
     direction of flow.

22.  The possible abrasive loss of anthracite coal media due to  air
     scour was evaluated  in a speedup test to simulate the total
     period of abrasive exposure which the coal might experience in
     a 20-year life.  The abrasive loss was found to be negligible.

          Conclusions Regarding Wastewater Filter Design

23.  The use of some form of air-scour auxiliary or some form of
     surface wash auxiliary is essential to the satisfactory func-
     tioning of wastewater filters comprised of deep beds (2 to
     5 ft) of granular material which are backwashed after several
     feet of head loss development.  The auxiliary and the backwash
     routine must be appropriate to the filter media.  For example,
     subfluidization wash is limited to single-media filters be-
     cause stratification is not essential (or even desired) for
     such filters.  Fluidization capability is essential for dual-
     or triple-media filters to permit restratification of the
     layers in their desired positions at the end of the backwash.
     Air scour and water  backwash simultaneously during overflow is
     primarily useful on coarse sand filters because finer media

-------
     will be lost due to the violence of the combined air and water
     action.  However, the simultaneous use of air and water can be
     useful on dual- and triple-media prior to the onset of backwash
     overflow.

     The above conclusion is not intended to apply to all types of
     wastewater filters such as the various proprietary filters
     with their special backwashing provisions.  Such filters and
     provisions were not studied in this research.

24.  The use of graded gravel to support the filter media without
     special provisions to prevent gravel upset is not recommended
     where the simultaneous flow of air scour and backwash water
     can pass through the gravel by intention, or by accident, due
     to the danger of moving the gravel and thus upsetting the de-
     sired size stratification of the gravel.

25.  Media-retaining underdrain strainers with openings of less
     than 1 mm are not recommended for wastewater filters, due to
     the danger of progressive clogging.

26.  The filter influent feedwater (e.g., secondary effluent) is
     not recommended as a backwash water source because of the
     danger of progressive clogging of underdrain strainers and/or
     gravel.  The advantages of using feedwater do not justify the
     risks that result therefrom.

27.  Air scour is compatible with  dual- or triple-media filters
     from the standpoint of minimal abrasive loss of the coal
     media.  However,  the backwash routine must be concluded with a
     period of fluidization and bed expansion  to restratify the
     media layers after the air scour.

28.  The coal and sand sizes for dual-media filters should be se-
     lected to encourage some intermixing at the interface to
     achieve improved filtrate quality.  However, to prevent some
     of the fine sand from reaching the top surface of the coal,
     the fines should be skimmed from the sand after installation
     of the sand in the filter and hydraulic grading by backwashing.
     This skimming is a desirable construction specification even
     though it may be inconvenient in the construction scheduling.

29.  In the filtration of secondary effluent,  a coarser surface
     filter media is favored to achieve longer filter run length
     and greater solids capture per unit of head loss development.
     Substantial difference was observed when comparing a strati-
     fied dual-media filter having a coal size of 1 to 2-mm with an
     unstratified coarse sand filter of 2 to 3.6-mm sand.  The 1 to
     2-mm coal gave much shorter run length.  As coarser coal sizes
     are considered, backwash rates required for fluidization may

-------
      become excessive (uneconomical),  and the advantage of sub-
      fluidization backwash becomes increasingly attractive.  How-
      ever, the choice of subfluidization backwash dictates the use
      of a single-media filter and an air-scour auxiliary rather
      than a surface-subsurface wash auxiliary.

30.   The use of coarser filter media is also favored by the desire
      to eliminate supporting gravel and fine slot strainers in the
      underdrain design.

-------
                      III.  RECOMMENDATIONS
The objectives of this research project have been largely fulfilled.
Some require additional work, and, as is the usual case, the re-
search conducted has led to the knowledge of additional work needed.
The following recommendations for additional work are offered;

1.  The work reported herein on the stability of double reverse
    graded gravel as a media support should be repeated using fil-
    tered water as a backwash source.

2.  The work reported herein on prediction of the expansion of coal
    filter media should be refined to improve its accuracy, perhaps
    by incorporation of a sphericity measure in the expansion model,

3.  The problem of media loss when backwashing with air and water
    simultaneously during overflow should be studied in a system-
    atic manner which includes the study of the prominent media ma-
    terials and sizes as well as a range of air and water flow
    rates.

4.  A comprehensive field investigation of existing full-scale
    wastewater filters should be made to see if the backwash rou-
    tines in use are maintaining the filters in good condition.
    The study should include the prominent proprietary filters cur-
    rently in use because of the unusual backwash routines employed
    in some of these filters.
                                10

-------
             IV.  BACKWASHING-POTABLE WATER EXPERIENCE
Rapid filters are washed to restore their capacity when the efflu-
ent quality becomes unacceptable, or when the pressure drop through
the filter reaches a predetermined value.  For gravity filters, the
terminal head loss selected is usually the actual head available;
thus operation beyond this time results in an inability to maintain
the desired filtration rate.  In some cases, filters are backwashed
on a regular time cycle based on experience with the two criteria
above, or at more frequent intervals dictated by past maintenance
problems created by excessively long filter runs.  Filter runs in
different plants may vary from 12 hr to several days, one day being
considered an acceptable average value.

The filter is usually washed by reversing the flow of water through
the filter.  In typical United States practice today (1975), the
rate is adequate to lift the grains of filter medium into suspen-
sion, that is, the rising water causes expansion of the filter me-
dium.  The deposited material is thus flushed up through the ex-
panded bed and to waste through the washwater gutters.

If the backwash is not effective, dirty filter problems such as
filter cracks and mud balls may occur.  Inadequate cleaning leaves
a thin layer of compressible dirt or floe around each grain of the
medium.  As pressure drop across the filter medium increases during
the subsequent filter run, the grains are squeezed together and
cracks form in the surface of the medium, usually along the walls
first.

The heavier deposits of solids near the surface of the media break
into pieces during the backwash.  These pieces, called mud balls,
may not disintegrate during the backwash.  If small enough and of
low enough density, they float on the surface of the fluidized
media.  If larger or heavier they may sink into the filter, to the
bottom, or to the sand-coal interface in dual media filters.  Ul-
timately they must be broken up or removed from the filter or they
reduce filtration effectiveness or cause shorter filter runs by
dissipating available head loss.

Early sand filters of the late nineteenth century in the United
States were provided with low washwater rates, 8 to 15 in. rise per
minute (in./min), which did not expand the bed, or expanded it only
slightly.  Auxiliary methods of agitation were provided in an at-
tempt to clean a filter.  These methods included mechanical rakes
that stirred the sand as they rotated in the filter sand and air
agitation systems in which air was introduced into the bottom of
the sand bed before or during the water backwash.  Neither method
was completely successful, due to inadequate washwater rates.
                                 11

-------
 United States potable water practice has  abandoned  these methods  in
 favor of a high velocity backwash (commonly 24  to 30 in./min  for  5
 to 10 min duration).   This  wash rate results in about 15 to 30%
 expansion of the sands currently specified,  depending upon water
 temperature, media gradation,  and specific  gravity.   Early work
 [62]  on the high velocity wash indicated  that with  the fine sands
 then  in use, 50% sand expansion gave the  best washing results.
 However, the trend to the use  of coarser  sands  results in exces-
 sively high wash rates if 50%  expansion is  to be achieved, so it  is
 seldom attempted in present practice.

 The high velocity wash commonly employed  in  the United States did
 not solve all problems with dirty filters,  and  it has created prob-
 lems  with the shifting of finer supporting  gravel layers when they
 are used.  The provision of a  surface wash  system which introduces
 high  velocity water jets before and  during  the  backwash has largely
 solved the problem of dirty filter medium for potable water filters,
 but has not solved the problem of shifting gravel.

 Evidence of the benefits of surface  wash  led to its wide adoption
 for potable water filters in the United States  [60,10].  Surface
 wash  is introduced at pressures  of 45 to  75  psig through orifices
 on a  fixed piping grid or on a rotating arm,  located  1 to 2 in.
 above the fixed bed.   Surface  wash flow rates are about 1 in./min
 for the rotary type,  and 3  to  6  in./min for  the fixed  nozzle  type  [3]
 The desired operating sequence involves draining the  filter to the
 wash  trough level or  below,  applying the  surface wash  flow with no
 concurrent backwash flow for 1 to  2 min to break up surface layers
 on top of the media,  then continuing the  surface wash  with concur-
 rent  backwash flow for several minutes until the backwash water be-
 gins  to clear up.  The concurrent  application may be  at two rates,
 a  low rate to barely  immerse the surface wash jets in  the media
 followed by a period  with normal bed expansion.  The  surface wash
 is then terminated and water fluidization backwash alone follows
 for 1 to 2 min to  stratify  the bed,  which is only important in dual-
media filters.

 British and  European  continental backwashing practice  continues to
use low rate  backwash with  little  or no bed  expansion  assisted by
 air scour.  This continued use has sparked a, renewed interest in
 and use of  such practice  in the  United States since about 1965.

The interest  in air scour has  also been stimulated by  the problem
of shifting gravel and  the more  difficult backwashing  of wastewater
 filters.  There is also  interest in  the use of underdrains with
 fine  strainers that do not require supporting gravel,  a system
which was  abandoned in  the early twentieth century due to clogging
and corrosion problems.

Air scour consists of  the distribution of air over the entire fil-
 ter area  at the bottom of the  filter media so that it  flows upward
                                 12

-------
through the media.  It is used in a number of fashions to improve
the effectiveness of backwashing, and/or to permit the use of lower
backwash water flow rates.  The air may be used prior to the water
backwash or concurrently with the water backwash.  When used con-
currently during backwash overflow, there is legitimate concern
over potential loss of filter media to the overflow due to the vio-
lent agitation created by the air scour.  When air is used alone,
the water level is lowered a few inches below the overflow level to
prevent loss of filter media during the air scour.

Air scour may be introduced to the filter through a pipe system
which is completely separate from the backwash water system, or it
may be through the use of a common system of nozzles (strainers)
which distribute both the air and water, either sequentially or
simultaneously.  In either method of distribution, if the air is
introduced below graded gravel supporting the filter media, there
is concern over the movement of the finer gravel by the air, or
especially by air and water used concurrently, by intention or
accident.  This concern has lead to the use in some filters of
media-retaining strainers which eliminate the need for graded sup-
port gravel in the filter.  However, these strainers may clog with
time, causing decreased backwash flow capability or, possibly,
structural failure of the underdrain system.

In view of the concerns expressed above and the renewed interest in
air scour in the United States, a summary of European air-scour
practice is appropriate because it has been used there since the
beginning of rapid filtration.

British potable water practice has included the use of air scour
for many years.  Air scour is used alone first, followed by water
backwash.  Plastic strainer nozzles with 3-mm slots are commonly
used in the underdrains and are generally covered by layers of
graded gravel to support the media.  Single-media sand filters have
been the most common, but dual-media filters are becoming more com-
mon since about 1970.  In the single-media sand filters, the wash
rate is only intended to reach minimum fluidization velocity with
only 1 to 27, expansion of the bed.  The sand grain size used in
Britain for potable water filtration is about the same size as in
the United States, although they specify the range of size of the
sand (e.g., 0.5 to 1 mm, meaning that the sand would pass a 1-mm
sieve and be retained on a 0.5-mm sieve) rather than the effective
size.  In the backwash operation, the air scour is intended to
loosen the dirt and is followed by the water backwash to flush away
the dirt.  Air is introduced through the gravel at rates of 1 to
1.5 scfm/sq ft, sometimes up to 2 scfm/sq ft, followed by water at
8 to 12 in./min  [118].  The upper rate is common in current prac-
tice.  These British water rates of flow are substantially lower
than the United States practice and are not sufficient for bed ex-
pansion.  The apparent success of such low rates must, therefore,
be attributed to the prior air scour and the use of single media
                                 13

-------
 which do not  require  restratification.   Problems with gravel move-
 ment have occurred but only in a few cases  [65].  The absence of
 such problems must be attributed to  the  low water and air  flow
 rates which are  presumably not sufficient to move the fine gravel,
 and  the  fact  that  air and  water are  not  used simultaneously.  Prob-
 lems with clogged  strainers have been controlled by using  strainers
 with 3-mm slots  after prior clogging problems were experienced
 using strainers  with  0.5-mm slots.

 In the British practice, the water level is lowered to  1 to 2 in.
 (5.75 cm) above  the fixed  bed  surface, a little below the  edge of
 the  washwater overflow which is  located  about 4 in. (10 to 15 cm)
 above the surface  of  the sand.   Then air is applied through mush-
 room type strainers that are used both for air and water distribu-
 tion.  The air scour  is applied  for  3 to 5 min and then is shut
 off.   Backwash water  is then injected immediately through  the same
 nozzles.

 The  backwashing  practice on the  European continent as described
 below is  obtained  from four sources  [61,37,70,108] and therefore
 probably  does not  reflect  the  diversity  of the continental practice.
 The  sources describe  several points  about continental practice
 which differ  substantially from  both  United States and British
 practice.

 a.    Deeper beds of coarser  sand  are used and the backwash
      of  these sands is  at  low  rates, with little or no expansion
      of  the bed  (<10%)

 b.    Backwash  is with air  and water simultaneously at low water
      rates  followed by  water alone to flush the solids out of the
      bed  and  to  the overflow.  The air rate is 2 to 4 cfm/sq ft and
      the water rate is  10  in./min for the smaller sand sizes (1 to
      2-mm size range);  for  the coarser sizes such as 2 to  3 mm or 2
      to 4 mm,  the rates are  6  to  8 cfm/sq ft air and 10 to 12 in./
     min water.  In fact,  the use of air and water simultaneously
      is considered absolutely essential because air alone compacts
      the bed  and causes solids to be pushed deeper into the bed
     between  the rising columns of air bubbles.   The potential
     danger of media loss  is acknowledged if the simultaneous air-
     water backwash is continued  during overflow.  It is suggested
     that if  loss is observed that the backwash water rate be re-
     duced during the simultaneous air-water backwash [37].  Mud
     balls are unknown  in Europe using this type of bed design and
     backwash system.

c.   Supporting gravel  is sometimes used but it is of the double
     reverse graded gravel arrangement,  i.e.,  coarse to fine to
     coarse in gradation [108].  Otherwise,  media-retaining strain-
     ers are  used but  the  dangers of  clogging and underdrain
     failure are acknowledged  [61],
                                 14

-------
The renewed use of air scour in the United States has been pat-
terned more after the British practice of using air scour alone
first, followed by water backwash.  United States air rates have
been typically 3 to 5 scfm/sq ft and the subsequent water wash is
above fluidization velocity to expand and restratify the dual-media
bed, typically 24 to 36 in./min.  The filters are usually equipped
with media-retaining underdrain strainers without graded gravel
support for the media.  Because of the fine sand media used in
potable water filters (0.5 mm effective size), the strainer open-
ings are very small (0.25 to 0.5 mm), and strainer clogging prob-
lems and some underdrain failures have occurred therefrom.  Because
of these problems, a reconsideration of the United States air-scour
design practice may be appropriate.
                                 15

-------
        V.  BACKWASHING WITH FLUIDIZATION AND EXPANSION
 In view of  the common practice  in  the United  States of backwashing at
 high velocity with  expansion  and fluidization of  the filter bed,
 studies were conducted on  this  method of backwashing.  Extensive lit-
 erature reviews have been  completed of both the chemical engineering
 fluidization literature  and the sanitary engineering backwashing lit-
 erature [4,5,28,80].  That literature will not be repeated in total
 in this Final Report, but  appropriate portions are included to sup-
 port the  stated conclusions.

                  Some Fluidization Fundamentals

 The phenomenon of fluidization  can best be visualized by passing a
 fluid  (gas  or liquid) upward  through a bed of solid particles in
 which  it  encounters a resistance to flow and  a resultant pressure
 drop Ap.  As the  flow rate V  is increased there is a linear relation-
 ship between Ap and V.   As V  is further increased a point is reached
 at which  the pressure drop is sufficient to bear the weight of the
 solid  particles.  Any further increase in flow rate causes the bed to
 expand and  accommodate the increased flow while maintaining the pres-
 sure drop Ap effectively the  same.  The fluidized bed thus formed
 closely resembles that of  a liquid [50].  The feature which distin-
 guishes the fluidized bed  from  other processes (fixed bed, filtration,
 etc.)  is  the motion of the particles within the bed.  The character-
 istics of an ideal fluidized bed and the distortions due to real con-
 ditions are indicated in Fig. 1.

 Within the  last two decades a flowering of thought has occurred in
 the field of fluidization, fertilized by the necessity of its use in
 the catalytic cracking of  heavy hydrocarbons into petroleum products.
 This has  given rise to five books  in English  [35,76,91,144,145] , six
 symposia, and countless  papers  in  the literature.  The above books
 and the reviews of recent work by Coulson and Richardson [33]  and
 Botterill [15,16]  form excellent general references for this study.

 Point  of Incipient Fluidization or Minimum Fluidizing Velocity -  Vfflf

This is the fluid velocity required for the onset of fluidization. It
could be defined exactly as point A in Fig. I for an ideal fluidized
bed.  For a real graded bed it  is defined by some as the intersection
of the two linear sections of the curve [35,76]  while alternate def-
 initions are presented by  others [110].

The bed is completely fluidized when the friction drag or pressure
drop across the bed is just enough to support the weight of the
 filter media [35].  Mathematically, this relationship is given by

     hpS = 4(ps - p)gd  -  e)                              (1)
                                 16

-------
    BED HEIGHT
                    REAL GRADED BED
                                        IDEAL UNISIZED BED
                   FIXED BED
     LOG
(PRESSURE DROP)
     Ap
                   'HYSTERESIS1
                     EFFECT
   FLU ID I ZED BED
DUE TO NONUNIFORM
FLUID DISTRIBUTION
(CHANNELING)
                                            IF PURE FLUID SECTIONS
                            PPAI rpAOFn RFn  PASS THROUGH BED
                            REAL GRADED BED  (SLUGGING)

                           IDEAL' UNISIZED BED

                             JVmf = MINIMUM FLUIDIZING VELOCITY
                        LOG (SUPERFICIAL FLUID VELOCITY)
          Fig.  1.  Characteristics of fluidized beds.
                               17

-------
 where

      h  = pressure drop across the  fluidized bed
      H   = height of expanded bed
      e   = porosity of expanded bed
      g  = acceleration due  to gravity
      p   = particle mass density,  and
      p   = fluid  mass density.

 The  simplest  bed expansion  can be worked  out by considering  a bed
 which is fluidized from initial porosity  e0  at height SLO  to  a poros-
 ity  e  and a height i,.   Since the volume of solids within  the bed re-
 mains constant,  then for a  bed of constant cross section  we have


     V1 " so)  = A(1  " e)'                               (2)

 Particulate or Homogeneous Fluidization

 In most  liquid fluidized beds  there is a  uniform increase of bed
 height for velocities  greater  than Vmf, and  the liquid passes smoothly
 and  appears uniformly  distributed within  the  interstices  of the solid
 particles as  shown in  Fig.  2.   This type  of  fluidization  with a uni-
 form distribution of particles was termed "particulate" by Wilhelm
 and  Kwauk [141].   The  condition of a filter  bed while being washed by
 water is  particulate.

 Aggregative or Nonhomogeneous  Fluidization

 For  most  gas-fluidized  systems, part of the  gas breaks through the
 bed  in the form  of bubbles which are considered shaped as a spherical
 cap  or a  sphere with a  collection of solid particles at the bottom
 as shown  in Fig.  2.  The  two-phase theory of  aggregative  fluidization
 postulates that  all gas  in excess of minimum  fluidization passes
 through the bed  as bubbles.  The bubbles  increase in size as they pass
 through the bed  and burst at the surface  of  the fluidized bed with a
 light scattering  of the  surface solid particles and those carried by
 the  bubbles.  This  is termed "aggregative" fluidization.  At higher
 rates of  gas flow, the  frontal diameters of the bubbles build to the
 diameter  of the containing apparatus, and a condition of the bed
 called "slugging"  is developed.  It cannot be overemphasized that all
discussion in this review referring to aggregative fluidization is a
description of a  two-phase system, composed of solids and gas.  Air
 scouring  in backwashing of filters is a three-phase system containing
 the media, water,  and air, as will be discussed later.

In recent years it has been shown theoretically [7,57,66]  as well as
experimentally [35,57,98,113]  that both particulate and aggregative
fluidization are  the ends of a continually changing spectrum of an
intrinsically unstable  system.  There are bubbles of liquid in a par-
ticulately fluidized bed, but  their magnitude is of the order of the
                                  18

-------




" " ».*•*" ~ ^ ."
***-*• '-'•l* *"+**'
* ',"\^»\\ ^r*%V»'
fTT'
V = V e













"-c^X^-
"> x.'"\'"- t i,
^ _ ,** x^»
M'f
V > V ,









V * >
*» ^
\ , *
X
** *
^ %•
*
* "" ^
•*• s "*k
s. %
w •«-
- ./
N ^
•**»**-*
\ * r
c-»l
                                 DILUTE PHASE
          PARTICULATE FLUIDIZATION
          V, V>»V, c-*l
IT1T IDT ITlf
BUBBLING SLUGGING DILUTE PHA
          AGGREGATIVE FLUIDIZATION
          (MOST GAS SYSTEMS)
Fig.  2.  Fundamental behavior patterns.
                    19

-------
 solid particles;  hence,  they are not discernible for a bed  of  small
 height.   However, their  effects are noticeable in beds of large  depth
 or when  the ratio of the density of solid to liquid  increases  [66],
 Thus steel or lead balls fluidized in water behave aggregatively
 while hollow resin or FCC catalysts fluidized by air behave particu-
 lately.

 It should be noted that  an aggregatively  fluidized condition cannot
 be maintained at  large average  porosities of the bed due  to an in-
 adequacy of available solid material to form the shells of  the bub-
 bles. Thus a reappearance of homogeneous fluidization can  be  ex-
 pected at very high mean bed voidages.  Several workers have pre-
 sented relationships for predicting whether particulate or  aggrega-
 tive fluidization will occur [35,104,141,145].

 Channeling

 Channeling [76,145]  is a condition in which the fluidizing  medium
 passes through a  bed of  particles  along a preferred  path.   Channeling
 can occur in both particulate and  aggregatively fluidized beds but
 is more  common and pronounced in gas-fluidized  beds.   In  an aggrega-
 tively gas-fluidized bed,  channeling can  be described  as  follows.

 The majority of the gas  passes  through in the form of  bubbles  ran-
 domly distributed throughout the bed.  When these  bubbles are  not
 randomly distributed but tend to rise through the  bed  along a  pre-
 ferred path,  then the bed  is approaching  a channeling  condition.
 Once this trouble has started,  it will lead to  a greater and greater
 degree of channeling.  Two  typical  cases  of channeling have been
 mentioned,  "through channeling," when the  flow  paths extend through
 the entire bed, and "intermediate channeling,"  when only a  portion of
 the bed  is subject to irregularity.  The design of the  gas  inlet de-
 vice at  the bottom of the bed has  an important  effect  on channeling.
 Channeling is  also affected  by  the  characteristics of  the solid phase
 such as  size,  shape,  and density.   It is more pronounced for finer
 particles  which tend  to  agglomerate.  Channeling tendencies are  al-
 ways smaller with porous plate  distributors than with multi-orifice
 distributors.

 Spouting

 Spouting  [12,76]   is  a  technique for  agitating a  bed of  particles too
 coarse to  fluidize well.  It  is in  a sense  a  combination of a dilute
 fluidized phase in the form  of  a rising spout surrounded by a down-
ward moving  fixed  bed  of solids.  Becker  [12] observed  that there is
 a maximum bed depth  above which spouting cannot  occur.  He  observed
 that  the minimum  spouting velocity,  expressed as a superficial veloc-
 ity,  required to  fluidize the maximum spoutable  bed depth is identi-
 cal with  the minimum  superficial gas velocity causing onset of aggre-
 gative fluidization for  the  same fluid and  solid system.  The author
differentiated between real  spouting and pseudospouting.  According to
                                  20

-------
him, pseudospouting is a case in which an internal channel is formed
under any condition.  Pseudospouting is described as a feeble form of
agitation.  It is nothing more than an empty vent containing essen-
tially no rising solids and owing its stability to the bridging tend-
encies of the particles.  In real spouting, agitation is caused by an
intense, continuous jet of gas piercing a quiescent bed and yielding
its energy to generate a well defined stream of particles, the spout,
whose upward motion is balanced by the downward motion of the sur-
rounding annulus.

Gas-Liquid Fluidization

Some research workers [39,89,119] have studied the use of gas-liquid
fluidized beds.  In gas-liquid fluidization, the liquid flows upwards
through a bed of solid particles which is fluidized by the flowing
liquid, while the gaseous phase moves as discrete bubbles through the
liquid fluidized bed.  Gas-liquid-particle operations, also called
three-phase systems, are of comparatively complicated physical nature.
Three phases are present, the flow patterns are extremely complex,
and exact mathematical models of the fluid flows and mass transport
in these operations probably cannot be developed at the present time.
Description of these systems will be based upon simplified concepts.
These processes are, however, distinguished by their high rate for
various purposes (e.g., heat transfer), good phase contact, and wide
ranges over which the process can be varied.

For air scouring to be identical to aggregative fluidization, the
water from the filter must be completely drained and sufficient air
supplied to fluidize the sand bed.  The rate of air supply required
to fluidize a sand bed of average size (0.8 mm) would be 80-100
cfm/sq ft [4]  as compared to the 3 to 5 cfm/sq ft commonly used for air
scour in backwashing.  It is clear that a true case of aggregative
fluidization is not possible in a normal air-scour operation in fil-
tration.  But there are similarities between three-phase fluidization
and aggregative fluidization.  Three-phase fluidization in filter
backwashing exists when air is applied to a bed which is fluidized
with water.  A three-phase system of backwashing rapid sand filters
such as this is discussed by Camp [26] and Simmonds [112].  Note
that there is a danger of loss of filter media in this operation un-
less the air scour is stopped before the water level overflows into
the wash troughs.  The more common method of air scouring is a special
case of a three-phase system where the water level is lowered to just
above the top surface of the sand and then the bed is air scoured at
the rate of 3 to 5 cfm/sq ft without the concurrent upward flow of water
and without fluidization of the bed.

It has been observed [89,119] that an increase of the gas flow rate
often causes a decrease of bed expansion for three-phase fluidization,
whereas bed expansion increases as the flow rate of the fluid medium
is increased for two-phase fluidization.  It was reported [89] that
this reduction in bed height or porosity is more marked in beds of
                                   21

-------
 small  particles  than in beds of large  particles.   A contraction of
 48% has  been observed in a highly expanded  bed  of  0.28-nnn ballotini.
 The reduction in bed height was explained by  the hypothesis  that  a
 portion  of  the liquid moved upward in  the wakes of the  gas bubbles  at
 a velocity  much  higher than the average  liquid  velocity.  The  liquid
 velocity in the  rest of the bed was consequently reduced below the
 average  liquid velocity,  and the expansion  of the  bed is reduced  cor-
 respondingly.  This  reduction in bed height is  more than the gas  vol-
 ume present at any instant,  thus causing a  net  reduction of  total bed
 volume.

 Rigby  et al.  [101] have investigated bubble properties  in three-phase
 systems.  They showed that gas  bubbles have a greater tendency to rise
 in  the center of the bed than near the walls.  This tendency was
 observed even when bubbling at the base  of  the  column was quite
 uniform.  As  result  of this tendency,  these authors concluded  that
 the most favorable gas distributor may not  be one  which distributes
 gas evenly  over  the  bed cross section, as has generally been assumed,
 but one  which introduces relatively more gas  near  the walls  to
 counteract  the natural tendency for bubbles to  rise in  the center
 of  the bed.

 Ostergaard  [90]  measured  the  rate  of growth of  gas  bubbles formed in
 a liquid-fluidized bed at  a  single  orifice  of 3.00-mm diameter for  gas
 flow rates varying from 9  to  63 cc/sec.  The  experiments were  carried out
 with tap water,  atmospheric  air,  and sand particles having an  average
 equivalent diameter  of 0.64 mm.  The equivalent diameter was defined
 as  the diameter  of a sphere with  the same average particle volume.
 The bubble frequency at the orifice was measured by an electrical re-
 sistance probe connected to an  oscilloscope.  The bubble frequency  at
 the bed  surface was  calculated  from  cinephotographs.  The measured
 rate of  coalescence  was markedly dependent  on bed porosity, having  a
 relatively high value  near the point of  incipient fluidization and de-
 creasing with  increasing liquid velocity and  bed porosity.  This  is in
 general  agreement with  the results of Rigby et al.  [101], as is the
 observation that the main change in bubble  frequency occurs within  a
 relatively short distance  from  the orifice.    The rate of coalescence
 did not vary significantly with gas  flow rate.  Observations of bubbles
 emerging through the bed surface show that bubble shape is markedly
 dependent on liquid velocity.  A bed near incipient fluidization  is
 characterized by a high viscosity, and an emerging bubble is of nearly
 spherical shape,  whereas a fluidized bed of high porosity is charac-
 terized by a viscosity not very much higher than that of water, so
 that an emerging bubble is of spherical cap shape.  The author [90]
 also observed that no  individual bubbles were observed at the orifice
at zero liquid velocity.  This was probably due to the formation of
gas channels in the  fixed bed.
                                   22

-------
              Predominance of Hydrod.yn.amic Forces
             in Cleaning by Water Fluidizatlon Alone

The basic hydrodynamic behavior of a fluidized bed is frequently for-
gotten in trying to explain the removal of impurities from the grain
of a filter sand during backwashing.  Fair and Geyer [45]  mention
that "substances adhering to the filter grains are dislodged ... by
the rubbing together of the suspended grains."  Babbitt et al. [8]
state that the "purpose of such expansion is to cause the sand grains
to rub against one another."

From purely theoretical grounds the suspension of particles in a ris-
ing stream of fluid is expected to require a field of flow around each
particle, thus negating the concept of a number of particles rubbing
together when fluidized.  Considerable direct evidence [105,106,135]
as well as most correlations [4,76,91] are based on the assumption that
the particles are uniformly distributed within the beds.  Thus, exper-
imental verification of these correlations implies that the assumption
of uniform distribution is reasonably valid.  Further qualitative sup-
port of the above assumption is the fact that particle attrition [145]
is negligibly small in fluidized beds and also the fact that however
well filters are backwashed, sand growth by layers of deposited mate-
rial frequently occurs in significant amounts.  Johnson and Cleasby
noted a growth from 0.43 mm to 0.65 mm in 14 years at the Ames plant
[67].

The most significant work which removed the above from the realm of
postulates to that of fact is Rowe's studies of "Drag Forces in Hy-
draulic Models of Fluidized Beds I, II" [105,106].  He showed in a
fundamental study that the drag forces on spheres arranged in regular
arrays is extremely sensitive to the separation between the particles.
The required modification to the drag coefficient for a single parti-
cle CD> due to neighboring particles, was effectively to multiply it
by [1 + (0.68/6)], where 6 = x/d, the dimensionless spacing of the
particles based on the particle diameter, d, and the clear distance
between particles, x.  Particles in rhombohedral packing were  found
to be subjected to a drag about 70 times greater than that on an iso-
lated particle for the same superficial velocity of the fluid  [106].
The value of the drag coefficient given by the above expression when
6 = 0.01 is 69 and was considered to refer to the maximum condition
[105].

Rowe's studies showed that small local changes of particle concentra-
tion were unstable because they required a very large change in the
velocity distribution.  A local decrease in particle concentration of
37,, required the velocities to be doubled.  It can be seen that as 6"*0,
the drag coefficient -» »; however, the expression does not apply for
6=0.  The studies indicate the existence of lateral repulsive hydro-
dynamic forces between particles; these became extremely large as the
spacing between particles was reduced.  Thus physical contacts between
particles in fluidized beds were extremely limited, and the particles
                                  23

-------
 were uniformly distributed in the fluid field.   Rowe clearly indicates
 that his development does not eliminate the existence of particle  con-
 tacts and concentration effects,  but only shows that they cannot per-
 sist and that their effect is negligible.  Adler and Happel [1] have
 also indicated that solid-solid frictional effects in the low porosity
 range of fluidized beds where they should be most significant were
 inconsequential.

 In a study of the stability of particulate fluidization,  Jackson de-
 veloped general equations of motion for a fluidized assembly of iden-
 tical particles [66] .   In discussing the equations of motion, he ne-
 glected terms due to the direct interaction by  collision and justified
 it in two ways.  First, if collisions between particles  were of com-
 parable importance to drag forces then the number of particles per
 unit volume would be expected to  decrease with  height above the bed
 support,  in the same way as the pressure of the atmosphere decreases
 with height above the earth's surface.   Second,  a posteriori justi-
 fication was  provided by the fact that the equations of  motion ob-
 tained by neglecting collisions gave a good qualitative  account of the
 main phenomena of fluidization.

 Murray [85] summarized Jackson's  arguments and  also gave  some of the
 other reasons for the  fact that negligible particle collisions occur
 in fluidized  beds.   He stated that

      Collision forces,  which are  a  form of particle pressure,
      are  also small,  since,  if such a term were  important  it
      would probably increase with rip  (ru,  = number density  of
      the  particles).   This would  result in a  gradation in  n«
      from the surface  into the bed  from zero  to  a finite val-
      ue,  this is  not observed.  The surface appears  to be  a
      discontinuity.  Furthermore, observation of  particle
      flow round a bubble by  X-ray techniques  seems  to  show
      little or no contact  interference  between neighboring
      particles.   Also,  if  collisions were  frequent,  the noise
      would be noticeable,  which is  not  the  case.

 Though Murray was chiefly  concerned with  aggregative  fluidization, it
 is  a  well  known fact [4,35,85,145]  that  the fluidized  section of the
 bed outside the bubbles is very similar  to  a  particulately fluidized
 bed at minimum fluidization.  The hydrodynamical  behavior  of  the con-
 tinuous phase  is hence  similar  to that of  a particulately  fluidized
 bed.

 In a  recent two-dimensional  study (which may  not  effectively  extra-
polate to  three dimensions) using a monolayer  of fluidized  particles,
Volpicelli et  al. [135] indicated the presence of  inhomogeneities in
 particle distribution within  the bed.  They presented  photographic
 stills from their motion picture,  study of  the fluidized beds.  A
 study of  these photographs by  the author of this  report showed that
particle contacts are rare, which confirms  Rowe's  studies.  Their
                                  24

-------
studies must be interpreted with caution since they used steel balls
and water as one of their systems; this system will exhibit marked
aggregative tendencies.  This study also indicated that particle flow
patterns switch from a circulation regime to a regime of quasirandom
motion as the voidage increases.  This characteristic, confirmed by
other workers too, has important implications in the development of
criteria for backwashing at optimum rates.

In a recent Russian theoretical study, on the pseudoturbulent diffusion
of particles in homogeneous suspensions using tensor analysis, turbu-
lence equations were developed for a two-phase system by Buevich and
Markov [20],  In a discussion on the .collisional dissipation of energy
the authors stated that particles undergoing collision have step
changes in velocity which will always be small for dilute suspensions.
They also said that the collisions of particles suspended in a liquid
are characteristically very gradual, and there are no step-wise
changes in the particle velocities.  The latter is due to significant
increases in the pressure in the liquid layer between the particles
as they approach each other and to the need for "squeezing out" this
layer before direct contact of the particles occurs.  An analogous
effect also occurred as particles approach a solid wall and in lubri-
cation processes, when the lubricating liquid in the gap between bear-
ing and slider played the part of this liquid layer.  Hence, they con-
cluded that any model of energy dissipation based on elastic colli-
sions will be in error by at least an order of magnitude.

Ruckenstein [107] developed a physical model for a homogeneous (par-
ticulate) fluidized bed, using the equation of motion of one particle
which is part of an ensemble of particles in interaction with a fluid.
The equation is established by neglecting the interaction by collision
of the particles of the ensemble.

All the above evidence in the fluidization literature pinpoints one
single fact:  the effect of collisional interactions between particles
in the fluidized state is comparatively insignificant.  This fact, now
becoming more and more accepted in the sanitary engineering literature
[4,6,26,132], indicates that the age-old argument whether abrasions
between particles or the hydrodynamic shear forces are the predominant
cleaning mechanism, should finally be laid to rest.  This conclusion
is also one of the basic assumptions of the theory of optimum back-
washing developed in this chapter.
                                   25

-------
          VI.   OPTIMUM CLEANING BY WATER BACKWASH ALONE

        Particulate Fluidization and Optimum Turbulence —

                   Evidence From the Literature
 Analyzing a filter being backwashed,  qualitatively, will  indicate that
 at minimum fluidization individual  particles have no motion  and  that
 frequently the  fluid  motion is  streamline,  and hence cleaning of the
 media will be negligible;  at the  other  extreme overexpansion of  the
 bed will  also reduce  the cleaning action due to  large  separations be-
 tween particles.   This  macroscopic  analysis indicates  that somewhere
 between the two extremes lies an  optimum condition which  we  seek.

 The review of the  literature quoted in  this section indicates a  strik-
 ing phenomenon  discovered  in particulate fluidization  research:  the
 existence of a  maximum  value for  most turbulence parameters  at a
 porosity  of 0.65 to 0.70.   This was the fact that originally led
 Amirtharajah toward inferring that  the  elusive condition  of  optimum
 backwash  would  probably be  centered around this porosity  [4].  This
 section summarizes his  theoretical  and  experimental proof of this
 hypothesis [5].

 Considerable evidence was collated  in a previous section  to  show that
 particle  abrasions or collision are inconsequential in a  fluidized
 bed.   This fact leads immediately to the deduction of  two very impor-
 tant hypotheses:   (1) the present United States mode of cleaning
 filters by fluidization  has  an intrinsic weakness in the  process it-
 self .  and (2) the  most  that  can be  achieved from the process is  to
 backwash  at  flow rates which will produce the maximum  turbulence and
 the maximum  shear  in the fluid-par tide field, for this is the prin-
 cipal  mode of cleaning.  The  first weakness is being remedied by the
 use of backwash auxiliaries.such  as air scour or surface washers.
 The second problem is far more tractable both theoretically  as well
 as  experimentally  with  the  systems we have at present  and the know-
 ledge  we  have in fluidization.

Hanratty,  Latinen  and Wilhelm [56] were the first to use Taylor's tur-
bulence equations  to describe the diffusion of a tracer dye  in par-
 ticulately fluidized beds.   They established the mixing parameters —
eddy diffusivity and the scales and intensities of turbulence.   The
experimental studies were made in a 5.40-cm Lucite tube, by  admitting
methylene  blue dye  from  a central location to the bed  of  fluidized
particles.  Four different systems consisting of glass spheres of dia-
meter  0.47, 0.93,  and 3 mm,  as well as silica spheres  of 1.84 mm were
used for  the solids in the bed.   In all runs except two,  a constant
expanded bed height of 20.3  cm was maintained, and the porosity was
 adjusted by changing the amount of  solids in the bed.  Two types of
 sampling  traverses — radial  and centerline — were used to measure the
 spreading  of the dye.
                                  26

-------
The results indicated that the theoretical equations of Taylor's
theory of turbulence were applicable for diffusion in a particulately
fluidized bed.  A minimum Peclet number, i.e., maximum eddy diffu-
sivity, was found for all particle sizes and corresponded to a poros-
ity of 0.70.  This maximum eddy diffusivity was directly related to
the maximum in the length scale of turbulence, which also occurred at
this porosity.  The intensity of turbulence increased continuously for
all porosities.  For the bed of 3-mm spheres the maximum scale of
turbulence was approximately 4 mm.

Hanratty et al. did not attempt to provide a quantitative explanation
for the minimum Peclet number at the porosity of 0.70.  Qualitatively
explaining the observed phenomenon in terms of the random-walk model
they stated

     Mixing in a packed bed was found to be explainable in terms
     of a random-walk model, and it is suggested that a dense
     fluidized bed retains elements of this mechanism.  The dis-
     tance a fluid element must side step in order to pass around
     a particle decreases as the bed is expanded.  Eventually, at
     a fraction void of 0.70, a fluid element may begin to flow
     past solid particles without the necessity at each level of
     flowing laterally in order to evade a particle.  Beyond the
     critical fraction void, in dilute beds, the turbulence is
     particle generated, and the eddy diffusivity is a direct
     function of particle population, leading to an increase in
     the Peclet number as the velocity is increased.

Cairns and Prausnitz [21,22] studied macroscopic mixing and longitu-
dinal mixing in solid liquid fluidized beds.  The studies in longitu-
dinal mixing were made by determining the electrical conductance
break-through curves using very small electrical conductivity probes
with a step-function input of salt-solution tracer.  The principal
advantage of the conductivity method is that it enables continuous
monitoring, and hence the tracing, of transient velocities in the
system.  Longitudinal eddy diffusivities were determined for 1.3- and
3.0-mm lead spheres and 3.2-mm glass spheres in 2- and 4-in. diameter
beds at a distance of five bed diameters from the injection point.

The analysis of the data was based on a statistical model developed by
H. A. Einstein in connection with the motion of pebbles in a water
stream.  The model gives an easy and rapid method of determining  the
Peclet groups from the experimental data.  The longitudinal eddy dif-
fusivities were determined for various  solids-to-column-diameter
ratios, various radial positions and various void fractions.  The re-
sults were consistent with the fact that the fluidized bed was con-
sidered as a transition between a packed bed and an open tube.  It
was found that the ratio of longitudinal to radial eddy diffusivity
was approximately 20 to 30.  Thus, the  rate of longitudinal mass
transfer was very much greater than the radial transfer.
                                  27

-------
 In all cases a maximum of the longitudinal eddy diffusivity occurred
 at a porosity of 0.65 to 0.70.   This maximum was much more  pronounced
 for the lead sphere system than for the glass sphere  system.   On
 defining the Peclet group in terms of the diameter of the solids  in
 the system,  a minimum Peclet number occurred at a porosity  of  0.7  in
 all cases.   However,  for the glass sphere system an asymptotic mini-
 mum value was reached for all porosities greater than or equal to
 a porosity of 0.7.  The authors concluded that the eddy diffusivity
 was strongly affected by the density and concentration of particles in
 fluidized beds and  a  maximum in the mixing properties occurred at  a
 fraction voids of 0.7.

 In the study reported later the authors  investigated  macroscopic mix-
 ing [22]  with a similar experimental setup.   They measured  the fre-
 quency distribution of  fluctuations and  the correlation coefficients
 at two points separated by a known distance.   From the mixing  data,
 radial eddy  diffusivities,  scales  of turbulence,  and  intensities of
 turbulence were measured.

 Since  mixing-length theory suggested that the mixing  process in
 fluidized beds was  analogous to molecular diffusion,  a similar ana-
 lytical model for a steady-state system  in cylindrical coordinates  was
 used for analysis of  the data.  The results  showed that the scale of
 turbulence maximized  at  a  porosity of 0.70.   A plot of the Peclet num-
 ber indicated a minimum corresponding to  this  same porosity.   Detailed
 visual records of the behavior  of  the solids  and  the  fluid were also
 noted.   The  most active  particle motion was  found  to  occur in  the
 range  of e = 0.70;  the motion of the particles  changed from a  circu-
 lation pattern to that  of  random motion.   This  change  in flow  pattern
 was  also noted by Volpicelli et al.  [135],  as  recorded before, in
 their  studies  of a monolayer of particles  in  the vertical plane.

Lemlich and  Caldas  [75]  studied the heat  transfer  characteristics from
 the  wall to  the  fluid within a  particulately  fluidized bed.   They
 used a bed of  glass spheres  fluidized by  water  and  found that  the heat
 transfer coefficient maximized  at  a  transition between two regimes  of
 flow.   The lower regime  indicated  limited  axial mixing, while  con-
 siderable mixing was evident in the temperature profiles at the higher
 rates  of  flow.   The maximum heat transfer  coefficient  occurred at
porosities of  0.66  and 0.81  for solid particles having diameters of
0.50 and 0.29 nun, respectively.

Handley  et al.  [55]  studied  the mechanics  of  fluid and particles in a
particulately  fluidized bed using unusual  experimental techniques.
They obtained  a  transparent  solid-liquid  system using  soda glass
 (density = 2.50, refractive  index = 1.52)  and methyl benzoate  (density
= 1.08; refractive index - 1.52).  An opaque white glass tracer parti-
cle having the  same properties  as  the transparent  soda glass was used,
and  the motion of this particle was studied by cine photography.   A
similar  technique was used in Ulug1s studies of flow visualization
 [132J.
                                   28

-------
Fifty histograms of displacement vectors measured from projected
stills showed a mean zero solids velocity and showed that standard
deviations were independent of radial or vertical positions in the
bed.  The mean zero solids velocity means that a single particle over
a sufficient length of time does not have a velocity, which is to be
expected because solids in a fluidized bed do not have a finite ve-
locity over a long period.  Thus, particle motion was random, and ho-
mogeneous although isotropic conditions were not obtained.  They then
applied Taylor's random-walk statistical analysis [49,125,126] to the
motion of the particles as well as the fluid regimes.  The results
showed that the root mean square (rms) of the turbulent fluid velocity,
passed through a maximum at some voidage between 0.44 and 0.75; the
extrapolated graph indicated a value near 0.70.  The rms of the tur-
bulent particle velocity similarly passed through a maximum at a void-
age between 0.44 and 0.75.  Using fluid dynamic pressure measurements
(made with a pitot meter), they also found that the vertical turbulent
fluid velocity component passed through a maximum at e = 0.68 to 0.70.
Unfortunately, due to the experimental limitation requiring trans-
parency, the authors did not investigate sufficient systems expanded
to porosities less than 0.65.

In one of the pioneering studies in fluidization, McCune and Wilhelm
[81] determined mass transfer characteristics by measuring the partial
dissolution of spherical and flake-shaped napthol particles in a flu-
idizing stream of water.  They found that the mass transfer character-
istics for 1/8-in. pellets maximized around a porosity of 0.70 to 0.75.

In a recent study, Galloway and Sage [51] , using an instrumented cop-
per sphere within a packed bed of spheres, studied the local thermal
transfer from the instrumented sphere.  Using the data of McCune and
Wilhelm [81] and Rowe [105,106], they established a boundary layer
model based on the behavior of thermal and material transfer from sin-
gle spheres and cylinders in turbulent fluid streams.  The studies
showed that a maximum mass transfer occurred at fraction voids of 0.70
for fixed beds.  Extending the use of the model for fluidized beds with
literature data, they showed that the maximum mass transfer correlated
with the maximum turbulence at the expanded porosity of 0.70.

Except for the qualitative observation of changing flow fields, the re-
searchers have not sought to explain why this maximum occurred at this
porosity.  Probably the answer to this lies in Jackson's studies [66].
Jackson probed the fundamental mechanics of the fluidized bed and con-
cluded that particulate and aggregative fluidization were both mechan-
istically unstable systems.  The instability manifested itself as
traveling waves of increasing amplitude.  His theory predicted that
disturbances similar to those of bubbles in aggregative systems would
also develop in particulately fluidized beds.  However, for particulate
systems these disturbances are of the same order of magnitude as the
size of the solids in the system and do not develop to a noticeable
extent, except in very deep beds.  The visual observations of Cairns
and Prausnitz [22] previously described were mentioned by Jackson as
                                   29

-------
 evidence  for  the  results  he  deduced  from his  stability study.  As  ad-
 ditional  evidence he  referred  to  the papers by  Slis  et al.  [115] and
 Kramers et  al.  [73].

 Kramers,  who  was  a  coauthor  of both  papers  [73,115],  studied in the
 second paper  [73] the longitudinal dispersion of  liquid in  a fluidized
 bed.  The study used  an experimental setup  similar to that  used by
 Cairns and  Prausnitz  [21], except that Kramers  et al.  used  very deep
 beds  (12  and  6  m) and fluidized glass spheres of  diameters  0.50 and
 1.0 mm [73].  In  order to avoid any  external  influences they took
 great care  to eliminate as far as possible  all  visible systematic
 eddies.   In fact  the  tubes used for  the  bed were  purchased  as a single
 piece and had no  connections or protuberances.

 The results indicated [73] a hump in the longitudinal diffusivity  at
 a porosity  of 0.7 for the 0.50-mm particles.  However,  at porosities
 greater than  0.75,  the value of the  diffusivity continued to increase.
 For the 1.0-mm  system the hump at the porosity  of 0.7 was barely per-
 ceptible.   Analyzing  these results in terms of  the reported maxima of
 Cairns and  Prausnitz's studies [21,22],  the authors  concluded that the
 eddy diffusivity was  composed  of  two parts.   One  part was supposed to
 be due to the eddies  produced  by  individual particles,  and  this con-
 tribution passed  through  a maximum at the porosity e =* 0.70.  The
 other part, which strongly increased at  higher  values  of porosity, was
 thought to  be connected with the  presence of  local porosity flucta-
 tions which were  seen to  travel upwards  through the bed.

 All the above studies  have considered a  solid phase  and a pure fluid-
 izing liquid.   In order for  the above results to  be directly appli-
 cable to  backwashing,  it  is  necessary to assess whether the presence
 of a number of  small  floe particles  in the liquid phase will affect
 the turbulence  parameters.   Precisely this question,  the effects of
 solids on turbulence  in a fluid, was  studied by Kada  and Hanratty  [71].
 Using the same  technique  as  that used by Hanratty et  al. [56] to study
 turbulent diffusion in fluidized beds, mentioned  previously in this
 report, Kada  and Hanratty studied the effect of solids concentrations
 of 0.13 to  2.5% by volume on the  turbulent dispersion  characteristics
 of a pure fluid.  They found that one of the chief variables affecting
 the system  was  the slip velocity, which  was the difference  between
 the particle  and fluid velocities in the direction of  flow.  For the
 systems studied the slip  velocity was equal to  that of the  free fall
 velocity.   It was found that glass particles having diameters of 0.10
 and 0.38 mm and concentrations of 1.5 and 1.7% by volume had no ef-
 fect at all on  the turbulent dispersion  characteristics of  the fluid.
One of the  principal conclusions of  the  study was that  for  systems
with small  slip velocities the effect of solid  concentrations up to
 1.5% had no effect at  all on the dispersion characteristics.  This
 study provides  the final  argument for applying  the results  of the
 studies reviewed in this  chapter  to  the  backwashing of  filters.
                                  30

-------
The evidence recorded above is impressive, since several studies have
confirmed the central result.  The impressiveness lies in the fact
that totally different experimental techniques used to study differ-
ent, but hydrodynamically related characteristics, namely, scales of
turbulence, eddy diffusivities, Peclet numbers, particle and fluid
motions, mass transfer and heat transfer effects, all yielded the
surprising fact of maximization at a porosity of approximately 0.65
to 0.70.  The very few studies which have sought to explain the reason
for this maximization of the turbulence parameters indicate that the
diffusivity is probably due to two factors:  the eddies associated
with the particles and the porosity fluctuations which travel up the
bed.

The essential conclusions of the preceding literature review can be
summarized as follows:

1.   Considerable evidence exists in the fluidization literature that
     particle collisions in the fluidized state are of negligible
     consequence compared to the hydrodynamic effects.  This fact is
     also being realized in the sanitary engineering field and as a
     corollary it implies that the use of water fluidization alone to
     clean a filter is inherently unsatisfactory, and various auxili-
     ary scour systems can be used to overcome this weakness.

2.   The fluidization literature abounds with evidence that the fluid
     and particle fields in particulate fluidization can be described
     by the statistical turbulence theories, even though the actual
     fluid velocities are not in the turbulent regime.  It has been
     found that most turbulence parameters have a maximum at an ex-
     panded porosity of 0.65 to 0.70.  It is hence hypothesized that,
     within the constraint that fluidization is not an excellent
     process for cleaning, the best cleaning that can be achieved is
     by expanding the bed to these porosities.

3.   Only qualitative and semiquantitative attempts have been made to
     unravel the reasons for the optimum in the turbulence parameters.
     These have been based on the assumption that dispersion is
     caused by (a) eddies around individual particles and (b) the
     movement of porosity fluctuations through the bed.

The theory developed for optimum backwashing in the next section of
this chapter is entirely new and original; however, it draws suste-
nance principally from the conclusions and results reported and sum-
marized in this literature review.

                  A New Theory of Optimum Backwashinp

                       by Water Fluidization Only

It has been shown in the previous sections that filter cleaning dur-
ing backwash is due to the turbulence of the fluidized bed and
                                  31

-------
 hydrodynamic shear forces.   The voluminous fluidization literature
 quoted has shown that turbulent diffusion is maximized at the poros-
 ity of 0.65 to 0.70.   Some  mathematical theories have sought  to iden-
 tify this maximum with the  eddies around particles and with the fluc-
 tuations  in porosity  which  travel up the fluidized bed.

 The following new theory shows that hydrodynamic shear forces in a
 fluidized bed reach a maximum at this porosity of 0.65 to 0.70 for
 most systems considered.  Even if the turbulence parameters cannot be
 directly  related to the  shear forces at the present time, on  the
 basis of  the discussion  in  the previous section the cleaning  in a
 backwashed filter bed necessarily reaches an optimum with the maximum
 shear.  Since shear and  turbulence parameters are inseparably related
 it  should be possible as  a  future extension of the theory to  obtain
 an  analytical model which will relate the maximum in the  hydrodynamic
 shear to  the maxima in the  turbulence parameters.  This is also pred-
 icated  on the fact that  the actual fluid velocities in a  fluidized
 bed are in the transitional regime and that only the interaction with
 the solids produces a system having behavior similar to that  of a
 turbulent field.

 Consider  a one-dimensional  analysis of the motion of an elemental
 volume  of fluid  as in Fig.  3.  Let p be pressure intensity, S shear
 stress, V  velocity,  and  Ax,  Ay,  and Az the dimensions of the cube.

 The work  done by  shearing stresses is irreversible and is dissipated
 as  heat.   This loss in energy corresponds  to what is frequently called
 head  loss or friction loss  in fluid flow problems.

 Let the power dissipated  by the  torque composed of the shear  forces
 due to  S,  by P^,

      P.. = torque  X angular  velocity
           (SAxAy)Az X
            dV
          S ——AxAyAz .
By definition
           dV
     S = p, -:—  where jj, is absolute viscosity,
Therefore,
                                  32

-------
 (pAyAz)
    (a) FORCES
                               SAxAy
            Az
                                        dV
                   Ax
                       V1
    (b) VELOCITIES
Fig. 3.   Shear forces on an elemental volume of fluid.
                        33

-------
 Let the hydraulic gradient in the z direction or the head loss per
 unit length be (dh/dz) .

 Hence the power dissipated by the element of height Az and area of
 cross section AxAy,  moving with a velocity V' , and having mass den-
 sity p is
      P2 =     AZ ' AxAyPg ' V'

 Since the power dissipated by the shear  forces  corresponds  to  the
 head loss, Eqs. (3)  and (4)  give
                    =  (^AxAyAzpgV,


 that  is,
                        1
                    IhM  2 .
                    lz>J                                             (5)

Now,  [(dh/dz)pgV]  is the power dissipated per unit volume,  say P/C,
and G is  the velocity or shear gradient defined  as (dV'/dz); hence
Eq. (5) reduces  to
             I

      G =  ye/                                                      (6)

Equation  (6) is  the familiar form of Eq.  (5), originally derived by
Camp  and  Stein [27] as  the general power  dissipation function in a
three-dimensional treatment.  The derivation of  Eq. (6) in condensed
form  is also presented  in Fair, Geyer, and Okun  [46].

For a fluidized  bed with a superficial velocity  V, where V1 = V/e,
the hydraulic gradient  (dh/dz) = i, and (j,/p is kinematic viscosity
   ,  Eq.  (5) becomes
                                                                   (7)

Equation (7) has previously been presented by Camp [25] for calcu-
lating the shear forces during filtration and backwashing.

Equation (5) will now be put in the form most useful for the follow-
ing development of the theory of optimum backwashing:


           dV    f   V /dh\l2
                                  34

-------
that is,

                   -.1
              "'-"l)2                                             (8)
where

     S = shear intensity

     V = superficial fluid velocity


     (-:— J = head loss gradient.


An important property of fluidized beds arises from the fact that
particles suspended in a fluid require that the frictional drag of
the fluid exactly counterbalance the pull of gravity.  In effect,
this leads to the requirement that the head loss across a fluidized
bed must equal the buoyant weight of the particles [Eq. (1)].  Two
of the earliest researchers to report this well known property were
Fair and Hatch [47].  In differential form this result is

     dhpg = dz(po - p)g(l - e)
                S

that is


     /dh\   (PS "
(S)
                p     (1 - e)                                      (9)
Equations (8) and (9) above and Eq. (10) below, form the principal
equations of the optimum backwashing theory.

Richardson and Zaki's equation, modified by Amirtharajah and Cleasby
[6] for graded and irregular particle systems, is


     V = Ken                                                      (10)

where

     K  = f(V ,ty,d/D ) = constant for a particular system
             S      t

     \|l  = sphericity

     d  = particle diameter

     D  = diameter of tube or bed

     V  = unhindered subsiding velocity of the particle
                                 35

-------
      n  = expansion coefficient in Richardson and Zaki's equation
           (presented below) .

 The coefficient n is a function of the flow regime and the dimensions
 of the apparatus but is constant for a particular system.  For the
 flow regimes of interest under filter backwashing conditions,
n
/4
I
45 + 18 ;
       D
                            for 1 < Re  < 200
                                      o
                                                                  (11)
                                                                  ^   '
 where
            PV d
      Re  = 	 =  Reynolds  number based  on unhindered  subsiding
             ^      velocity.

 It  should be pointed  out  that Eq.  (11) was  developed  from fluidiza-
 tion studies of spherical particles and is  not  directly applicable to
 nonspherical particles  as Amirtharajah and  Cleasby had  presumed [6].
 Nevertheless, K and n are constants for a particular  filter media.

 Substituting Eqs.  (9) and (10)  in  Eq.  (8) gives
 S =
            n
                                     12
                                  -e)J
 that  is,
where
               (p.  -
  -a  (e11-1 - en)
                    n2
                                                                 (12)
     r            i
tt =  UsK  (PS " P)
                          = constant for a particular system.

The above equation is the basic equation of the writer's theory.  It
is a relation between the shear stress and the porosity in a fluidized
bed.

Let us analyze this function by classical optimization techniques to
determine the stationary points of the function as the porosity e
changes.
                                   36

-------
                                                    2
The simplest analysis is to consider  the function  S :

      O    O  / _ 1    _v
       1,



     i  < 0 when e  =
     de
Thus  the  stationary point is a maximum.   Alternatively, the following
simpler analysis  gives  the same result.
                                   37

-------
 Consider the  sign of  dS/de  as  it  passes  through the stationary point.
 From Eq. (13) ,

      dS    _   f        (n -  1)
      3—  > 0,  for  e <  s      '
      de                  n

      dS
      de  < 0,  for  e >     ~
                         n

Hence  the  stationary  point  is  a maximum.

For  a  typical  filter  sand with an effective  size of 0.45 mm  and  a
uniformity coefficient of 1.47, n for  the  top  3 in. of  the graded
sand is  3.54 at  25 °C.  For a uniform sand of 0.66-mm size, n = 3.2
at the same temperature  by  Eq. (11).

For  the  graded sand,  maximum shear  stress  S  occurs at

     e - (° -  1) =  (3.54 -  1)  =
     *>—        ~~    A  f. /     — I/ • / ^ «
            n         3.54

For  the  uniform  sand,

         ? 7
     e.^f-0.69

Thus, a  maximum  shear stress S occurs  in a fluidized bed at  the  po-
rosity e = (n -  l)/n, which corresponds to porosities of 0.69 to 0.72
for  real sand systems.   This is the main result of the  optimum back-
washing  theory.

The  above  result, derived entirely  from the  theory, indicates that
optimum  cleaning of the  filter by maximum hydrodynamic  shear forces
occurs at  the porosity of about 0.70.  This  theory, in  combination
with the literature cited in fluidization, which reviewed several
experimental studies  indicating an optimum diffusion at the porosity
0.65 to  0.70, provides an excellent theoretical framework for experi-
mentally studying optimum backwashing.  Detailed experimental studies
which provide confirmation  of  this theory  are presented in the next
section of this report.

The above theory is developed  from three equations which are valid
for all  types of flow regimes  in fluidization.  Camp and Stein's
equation is valid for viscous  as well as turbulent flows since it
only equates the energy, dissipated by shear  to the head loss.  The
constant head loss equation is valid for all fluidized  beds, and
Amirtharajah and Cleasby's  equation is a modified form  of a power
function of porosity which  is valid for all  shapes and  sizes of  par-
ticles.  Hence the theory developed is applicable for all particu-
lately fluidized systems in all regimes of flow.
                                  38

-------
As a fitting closure to the above theory, it is necessary to antici-
pate the results that can be derived by applying this theory to back-
washing in practice.  The theory predicts an optimum in cleaning at
a porosity of 0.70.  Consider a uniform sand bed of depth Jlo with a
fixed bed porosity of 0.43.  For the porosity to ,be 0.70 in the ex-
panded state of depth SL,

     SL  (1  -  0.70)  = jeQ (1 - 0.43);

therefore,

     Si = 1.9 A  .
              o

Hence the expansion required is about 90%.  This can rarely be a-
chieved in practice.  However, for a graded system the particle diam-
eters at the top of the bed are a fraction of the diameters in the
deeper sections.  Thus an expansion much smaller than 100% will cause
the porosities to be 0.70 in the top layers.  Since these layers are
the ones that remove most of the suspended matter in filtration, it
can be rationally expected that optimum cleaning of the top layers
will produce the best cleaning for the system.  Expansions higher
than that producing the porosity of 0.70 in the top layers will tend
to increase the porosities of the layers on top but will simultane-
ously cause the lower layers to reach the optimum porosity of 0.70,
hence we would expect only a negligibly small decrease in the opti-
mum cleaning.  This would cause a nearly asymptotic curve of optimum
cleaning to be produced for graded systems.

              Experimental Support for the Optimum Theory

Experimental Apparatus

A schematic layout of the experimental apparatus is shown in Fig. 4.
The arrows in Fig. 4 trace the path of water from the tap supply to
the outlet drain for filter F3 during a filtration run.  The main
pilot plant consisted of the university tap water supply (hot and
cold) blended in a thermostatically controlled mixing valve A
(Lawler Automatic Controls, Inc., Mt. Vernon, New York).  The blended
water passed through a centrifugal pump B, used as an in-line booster.
After being metered in the flowmeter C, the water passed through a
dual outlet; one end of this outlet fed the supply water to the mix-
ing tank D, while the other end provided the backwash water supply.
Each of these outlets was used singly, and the pump B was only used
during high rates of backwash since the normal tap pressure was
sufficient for most uses.

The influent to the filters was mixed in tank D with the chemicals
being added from constant head capillary feeders.  The influent water
was pumped from the mixing tank by a centrifugal pump E, which was
driven by a variable speed DC motor.  This enabled influent control
                                  39

-------
                     BACKWASH
TAP WATER SUPPLY      DRAIN —
    HOT COLD
                               I     -I     •!
                               i       i
  CHEMICAL
  FEEDER
                     INFLUENT
                     LINES
                  BACKWASH
                  LINE
                  MAIN
                1 INFLUENT
                  LINE
                   OVERFLOW
F3
     MIXING TANK
F2
       *—_
       kl
              r— SIPHON BACK-
                 WASH LINE

               TYPICAL
               PIEZOMETER
               CONNECTION
Fl
                                                 :   T
                  DRIP
                  SAMPLER
                                                FLOWMETERS
                                            H,
                                                EFFLUENT
                                                RATE
                                                CONTROLLERS
                                               NEEDLE VALVES
                                  TO DRAIN
          Fig. 4.  Schematic layout of experimental apparatus.
                               40

-------
to be achieved.  The main influent line trifurcated to the filters
via the filter valve system.  The effluent from each of the filters
Fl, F2, and F3 passed through its own flowmeter and then discharged
freely into a float operated effluent rate controller.  Eleven pi-
ezometer connections enabled the head losses to be determined at
every 3-in. depth of the filter.  The expanded height of the filters
was determined by a scale placed along the side of the filter.

For backwashing the filters, blended tap water was used and metered
in flowmeter C.  The backwash line passed via the valve system and
discharged in a drain.

The filters consisted of 6-in. inner diameter, 1/2-in. thick plexi-
glass tubes 4 ft 5 in. deep with a 3-in. high calming section at the
bottom.  The water was fed through 59 orifices of 1/16-in. diameter
in a 1-in. thick underdrain plexiglass plate.  Sets of orifices were
staggered from one another so as to provide a uniform matrix of ori-
fices on the entire plate.  The calming section was filled with
1/2-in. diameter glass marbles.  Two pressure gauges, one attached
to the center of the calming section and the other to the cover plate
of the filter, were provided to serve as safety gauges to warn of
dangerous pressures.

The solid particles composing the bed were supported on two stainless
steel meshes (No. 50 over No. 10) placed above the 1-in. plexiglass
plate with the orifices.  Sixteen pressure taps were located on two
rows on diametrically opposite sides of the filter to permit obser-
vation.

A mixing tank 3-ft high and 3 ft 6 in. in diameter equipped with a
slow speed paddle was used to provide the necessary detention time
for completion of the iron precipitation reaction.  Allowing 6 in.
of freeboard, this 180-gal. tank provided a 30-min theoretical deten-
tion time at a pumping rate of 6 gpm.

A 1.5-in. outlet diameter, self-priming centrifugal pump  (Teel Self-
Priming Centrifugal Pump, Model 1P746, Dayton Electric Manufacturing
Co., Chicago, Illinois) was used to pump the influent water from the
mixing tank to the filters.  The pump was driven by a 2 by 3.5-in.
pulley drive from a 1/3-HP DC motor equipped with a variable speed
control.(Westinghouse Hi-Torque Speed Control with DC Motor, Westing-
house Electric Corporation, Springfield, Massachusetts).  This system
enabled the influent pumping rate to be increased simultaneously on
all three filters by a gradual increase in the speed of the centrif-
ugal pump.  Thus influent control was achieved; any changes in flow-
rate affected all three filters to the identical extent.

The total flow to the mixing tank during filtration and the total
backwash rates were metered by a rotameter C which had a  range up  to
13 gpm.
                                  41

-------
 The effluent  from each filter passed through a rotameter  suitable  for
 water  flow measurement from 0.2 to 2.0 gpm.   These flowmeters,  Gi,
 G£, and  63, were  calibrated by determining the time to  fill  a  liter-
 measuring cylinder.

 The effluent  from each filter passed via each flowmeter into a float-
 operated,  rate-of-flow controller.  The controller maintained  a con-
 stant  rate of filtration by holding a constant head on  a  needle valve
 outlet.   As the head  loss through  the filter increased  during  a run
 the float  valve gradually opened to maintain a constant filtration
 rate.  These  rate-of-flow controllers functioned  remarkably  well.

 Granular filter sand  (Fine sand, from Northern Gravel Co., Muscatine,
 Iowa)  was  used for this study.   The original sand had an  effective
 size of  0.455 mm,  a uniformity coefficient of 1.52, a specific  grav-
 ity of 2.648  and  a porosity of 0.412.   The graded sand  used  in  the
 latter stages of  the  study was  this original sand.   The uniform sand
 used in  the earlier stages of this study was prepared by  sieving in a
 Gilman set of United  States Standard sieves  on a  mechanical  shaker
 for 10 min.   The  uniform sand used was that  sand,  100%  of which
 passed sieve  no.  30 and was retained on sieve no.  35.   This  sand had
 an  arithmetic mean size of 0.548 mm or a geometric  mean size of 0.545
 mm  on  the  basis of the adjacent sieve  openings.  .Special  precautions
 were taken to insure  that the sand was identical  in all three  filters.

 Laboratory Analytical  Procedures

 The influent  suspension was prepared by dripping  a  stock  solution of
 ferrous  sulphate  (FeS04'7H20)  from a constant head  capillary feeder
 to  the mixing tank, into which  a metered quantity of water (4.5  gpm)
 was added  continuously.   This  flow rate was  sufficient  to apply  to
 the three  filters  at 7  gpm/sq  ft,  and  allowed a continuous bleed from
 the bottom of the mixing tank  as well  as continuous overflow.  The
 stock  solution had an  iron  concentration of  approximately 0.2 M in an
 acid solution of  strength  0.1 N to prevent precipitation  in  the
 bottle.

 Table  1  gives  a typical  analysis of  the University  tap water used in
 the  studies.  This is  a  hard well  water of high alkalinity and  total
 dissolved  solids and of  relatively constant  quality.  When ferrous
 sulphate in acid solution was added  to  this water a yellowish brown
precipitate formed.  A  sample of the suspension from the  influent to
 the filters,  after the normal detention in the mixing tank, was  fil-
 tered  through  a 0.45-|o,m  millipore  filter, and  the filtrate was  ana-
 lyzed  for  any dissolved  iron.   It was  found  that within the  accuracy
of  the equipment used no dissolved  iron was detectable.   This indi-
cated  that the precipitation was complete in  the mixing tank.  It
 should be noted that no  aeration or  addition  of air was required for
 the formation of the precipitate.  All  the runs were made without
 the addition  of any air  except  that which occurred  at the water  sur-
 face of the tank which was  open to the  atmosphere.
                                  42

-------
        Table 1.   Analysis  of University tap water.


          Characteristic of water         Concentration, mg/1


        Total dissolved solids                  680
        Total hardness as CaC03                 365
        Calcium hardness as CaC03               254
        Magnesium hardness  as CaCOj             HI
        Total alkalinity as CaCOs               270

        Calcium as Ca"*"1"                         102
        Magnesium as Mg**                        27
        Bicarbonate as HC03~                    330
        Chlorides as Cl~_                        17.5
        Sulphates as 804""                       160
        Fluorides as F~                           0.9
        Manganese as Mn                           0.0
        Iron as Fe"^                              0.03


The actual nature of the precipitate formed by addition of ferrous
sulphate to high alkalinity water has been the subject of considera-
ble controversy during the  last decade [30,52,114].   It has been
supposed by various workers that the precipitate could be ferric
hydroxide, ferrous carbonate, or a combined precipitate of both.  It
is suggested as a future project that wet chemical analyses be per-
formed on the precipitate to identify its character more definitely.

The main series of experiments consisted of 18 runs, 12 made on uni-
form sand filters and 6 on graded sand filters.  The influent and
effluent qualities were evaluated on the basis of iron content.
During these 18 runs nearly 5400 analyses for iron were made; thus
a simple and accurate method of analysis was required.

During the initial series  (run 1 to run 6) the standard method in the
water supply field, namely the 1, 10-phenanthroline method was chosen
[117].  The procedure was considerably simplified by using the pat-
ented single-powder formulation  (Hach Chemical Company, Ames, Iowa  -
FerroVer) which dissolves  and reduces the iron without any heating.
Since very few interfering ions were present the results were un-
affected by  such  interferences.  The developed color was observed at
510 nm on a  Beckman Model  B  spectrophotometer and the iron content
read on a calibration curve.

The above procedure  for measuring  the effluent iron had the following
weaknesses:   (1)  The  procedure  for measuring the quantity of reagent
using a scoop had possible errors  from  analysis to analysis;  (2)  the
transmittances of the final  effluent at  the 12-in. depth was too
high; (3) the molar  absorptivity of phenanthroline limited the possi-
ble  accuracy of  small changes in iron concentration.
                                 43

-------
 In order to alleviate these weaknesses the following modified  proce-
 dure was used in all the runs after run 6, and it improved  the accu-
 racy of the analyses considerably.   A new patented reagent  (Hach
 Chemical Company, Ames,  Iowa), disodium salt of 3-(2-pyridyl)-5,6-
 bis(4-phenyl sulfonic acid)-l,2,4 triazine, hereafter called Ferro-
 Zine, having a molar absorptivity of 27,900 was used.  This compound
 reacts with divalent iron to form a stable magenta complex  species
 which is very soluble in water and may be used for the spectrophoto-
 metric determination of  iron.  The absorption spectrum of the  complex
 has a sharp peak at  a wavelength of 560 nm and is uniform in develop-
 ment over the pH range from 4 to 10.  Further details of interference
 studies and statistical  data on multiple laboratory studies of Ferro-
 Zine can be seen in  Stookey [120].

 The above reagent as a single-solution formulation,  FerroZine  Solu-
 tion 1 (Hach Chemical Company,  Ames, Iowa),  can be added directly to
 iron hydroxides or carbonates for reduction and dissolution.   The
 procedure for analysis was to add 0.5 ml of FerroZine Solution 1 to
 25  ml of water and read  the transmittance on a spectrophotometer
 after allowing 5 min  for development of color.

 In  order to obtain increased accuracy of the transmittance readings on
 the spectrophotometer, a dual standard procedure,  based on the method
 of  ultimate precision as described  in Ewing [44]  was used.  This
 precision calibration method nullified all the stated weaknesses of
 the  FerroVer method considerably and increased  the  accuracy of the
 effluent quality analyses in the second and third series of runs.

 During the runs,  the influent suspension was sampled every half hour.
 These samples had iron concentrations of 7 mg/1,  and hence they had
 to  be diluted for obtaining readings within the  range of the calibra-
 tion curves.

 All the  samples  of influent and  effluent were  collected every  half
 hour.  Also,  the  initial effluent quality was  analyzed  at frequencies
 of  nearly a minute during the first  10 min  of a run to study  the
 initial  degradation  and  improvement  of effluent  quality.  These  sam-
 ples were also  analyzed  by the FerroZine test.

 While backwashing the  filters  after  a dirtying run,  samples of  the
 backwash  water were  collected periodically.  These samples had  con-
 siderable amounts  of  suspended iron  floe,  some as high  as 800 mg/1.
 These  samples were also  analyzed using  the phenanthroline procedure,
 the  samples being diluted  to  obtain  reasonable readings on the  spec-
 trophotometer.

During the  course  of the  research project  it was  realized that  addi-
 tional evidence of the effectiveness  of  backwash could be obtained by
 analyzing the  amount of  iron  left as  a coating on the sand after the
wash.  Not  only would this provide comparative evidence for studying
                                  44

-------
various expansions, but it would also prove beyond any shadow of a
doubt that abrasion in the fluidized bed was negligible.  This  of
course, was already anticipated from theory and other experimental
results quoted in the earlier chapters.

It was decided to evaluate two methods of washing the sand:  (1) a
physical wash and (2) a chemical acid wash.  The physical wash pro-
cedure was selected as outlined below.

     The filter was fluidized, and a long-handled scoop was used
     to remove a sample of sand from the top layers of the fluid-
     ized bed.  Each sample of sand weighed approximately 25 g.
     This sand was washed from the scoop into a 250-ml beaker
     using exactly 100 ml of distilled water.  The beaker had al-
     ready been weighed, empty and dry.  It was now weighed con-
     taining the sand, the water drawn by the scoop and the dis-
     tilled water.  A 1.25-in. magnet was placed in the beaker
     with the sand and water, and the sand was washed by the
     magnetic stirrer for 10 min  at a fixed speed.  It was found
     that the distilled water turned quite dark and cloudy and
     that considerable amounts of iron had been removed from the
     sand by abrasion between sand particles as well as by abra-
     sion with the magnet.  The supernatant iron suspension was
     stirred by the tip of a pipette and 25 ml were withdrawn and
     delivered into a 500-ml volumetric flask.  Distilled water
     was added to make up to 500 ml and this diluted solution
     (1:20) was analyzed for iron concentration using the Ferro-
     Zine standard calibration technique.

     The beakers containing the sand and the iron suspension
     (approximately 75 ml) were placed in an oven.  The water
     in the beakers was evaporated, and the beakers containing
     the dry sand vere cooled to room temperature and weighed
     again.

From the above readings the amount of iron removed from the sand  in
mg/g can be determined from the following formula:

Iron removed (mg/g) - [Cone, of iron in diluted solution  (mg/1)]

                      [Dilution Factor] x [Weight of water  (g)l
                    x       1000 x [Weight of sand (g)]           (15)

Note that the weight of water in grams is assumed to represent  the
volume of water in milliliters, and it is the total water in  the
beaker including whatever water is drawn with the sand  from the flu-
idized bed.  This formula gives the iron removed  from the sand  quite
accurately.

The above was  the procedure used  in all  the  analyses run on samples
of the graded  sand during  the  third series,  runs  20  to  25.  At  the
                                  45

-------
 end of each sand analysis, and prior to the beginning of the subse-
 quent run, the withdrawn samples of sand were returned to their re-
 spective filters.  Thus, the sand in all the runs was identical and
 no sand losses were allowed to develop.

 Chronology and General Descriptions of Filter Runs

 Twenty-five filter runs, of which 18 were dual runs,  were made.  The
 dual runs consisted of a dirtying run designated A,  the backwash of
 A, and a filtration run designated B.  In the following, a reference
 to a run means the complete dual run.  Runs A and B  are designated
 as such.  The first 12 runs (series 1:  1A, IB to 6A, 6B; series 2;
 7A, 7B to 12A, 12B) were made on filters with 12-in.  depth uniform
 sands of 0.548-mm mean size.  The runs in series 1 and 2 were  made
 at 7 gpm/sq ft, and each run (A or B) lasted for approximately 5 hr.
 Runs A or B had to be terminated due to the fact that the head losses
 developed were nearly 8 to 9 ft, and this was the maximum differen-
 tial height that could be measured on the piezometer  boards.

 The reagent used for iron analysis in series 1 was FerroVer,   In the
 first series of runs, the dirtying runs A were made on the first day,
 and the backwash and the filtration  runs B were made on the follow-
 ing day.  Since it was possible that some physical and chemical
 changes could have occurred in the solids removed, due to overnight
 standing, and also because in actual treatment plants backwashing is
 performed soon after a filter is removed from service, the two runs
 A and B, and the backwash of runs A in series 2 were  made on the same
 day.   Each dual run including backwash required about 15 hr. and two
 experimenters were required to work continuously to take the readings
 and make the analyses of iron.  In run 1 the samples  of water  at
 depths other than the full depth of 12 in.  were collected by a single
 experimenter and kept for about two to three days before all the
 analyses were completed.   It was found that iron in suspension tended
 to be deposited on the sides of the plastic sampling  bottles and that
 the effluent quality at 9-in. depth of the  filter measured on  a later
 date  was better than the effluent quality measured at the 12-in.
 depth on the day the run was made.   These results were invalid and
 were  not used in the analyses.  All iron analyses from run 2 onward
 were  made within a couple of hours  after sampling, and rational
 readings were obtained.   This could only be accomplished because  two
 experimenters worked.   The reagent  used  for series 2  and 3 was Ferro-
 Zine.

 The last six dual  runs  (run 20A,  20B through run 25A,  25B)  were made
 on 18-in.  depth graded  sands  with an effective  size of 0.45 mm and a
 uniformity coefficient  of 1.47.   This  set of runs was  called series 3.
These  runs were made using identical procedures  to those of series 2.
However,  in  addition  to  ferrous  sulphate, a nonionic  polyelectrolyte
 (Dow Chemical Company, Midland, Michigan -  Separan) was  also added to
 the influent  suspension  to  obtain a concentration of  0.10 mg/1 of
                                 46

-------
polyelectrolyte in the feed to the filters.  The polyelectrolyte was
added in the hope of magnifying the differences of backwashing at
different expansions.

The parameters used to study the effectiveness of backwash in all
three series were (1) initial effluent quality, (2) head loss in-
crease, (3) cumulative effluent quality in the run following back-
wash, (4) the backwash water quality during backwash, and (5) the
backwash water volume.

For series 3 the additional parameter based on the iron which is phys-
ically removable from samples of the backwashed sand was also used.
The variation of each of these parameters with porosity was studied
at six different porosities, 0.55, 0.60, 0.65, 0.70, 0.75, and 0.78,
on each of the filters.  The expansions needed to obtain these poros-
ities for the uniform sand were approximately 33, 50, 70, 100, 140,
and 190% respectively.  Since each run was made on a bank of three
filters and each series consisted of six runs, the effective number
of points for each series was 18.  The experiment was designed such
that each filter was studied at the six different porosities during a
series.  Thus by considering all 18 readings of a series of six runs,
the small variations between filters and the small variations from
run to run were averaged out.  Table 2 illustrates the format of the
experimental design for series 2, and how the backwash at the differ-
ent expansions was studied for the three filters Fl, F2, and F3.  A
similar format was used for series 1.

Table 2.  Experimental design for series 2.


                            Porosity during backwash
             0.55      0.60
K.un
number
7
8
9
10
11
12


F2
Fl
F3



Fl


F2

F3
Filter

Fl


F3
F2
number
F2

F3

Fl



F3

Fl
F2


F3

F2


Fl
 For the series  3,  the  format  of  runs  7  and 8,  as  shown in Table  2  for
 series 2,  was repeated three  times.   This procedure was adopted  to
 enable a particular run to be studied with the backwash expansions of
 30, 50, and 75% or 15, 40, and 60%.   This format  avoided two adjacent
 expansions such as 40  and 50% being  studied in a  single run, and the
 differences for purposes of comparison  were enhanced.   This was
 thought advantageous because  of  the  anticipated smaller differences
 in effectiveness of backwash  at  different porosities for graded
                                   47

-------
 systems.   The cause for the anticipated smaller differences  in back-
 wash for  graded systems has been discussed previously.

 The  expansions needed to obtain porosities of 0.58 to 0.82 in the  top
 layers  of the graded sand ranged from 15 to 75%,  and  these were much
 less than the expansions required for the uniform sand.  All  back-
 washing expansions  were controlled on the basis of the expanded
 heights of the fluidized bed during backwash.

 In order  to reproduce identical conditions with a clean  filter at
 the  beginning of  each dual run,  the following standardized procedure
 was  used  for all  runs.   The filter was expanded to the anticipated
 optimum at a porosity of 0.70 and washed for 15 min.  During  the runs
 1 to 10 the filters were washed  twice before each run began,  once at
 the  completion of the previous run and again immediately preceding
 the  start of a run.  However,  from run 11 onwards,  in order to achieve
 identical conditions to those  in a treatment plant, the  solids re-
 moved in  run B of the preceding  run were not washed until the follow-
 ing  run A.   Immediately before the dirtying run A,  the filter was ex-
 panded  to the anticipated optimum at a porosity of  0.70  and washed
 for  15  min.

 In the  case of series 1 and  2  the time of  backwash was the same for
 all  expansions; for the graded sand  studies  of  series 3, however, the
 operation was modified  slightly.   In series  3,  the backwash at dif-
 ferent  expansions was done  for different times, so that  the same
 volume  of washwater,  approximately 36  gal.,  was used during the total
 sequence  of washing (i.e., valve  opening, washing, and valve  closing).
 Also, while the bed was  fluidized, one experimenter collected  two
 samples of  the backwashed  sand at  different  times.  The  samples were
 collected after approximately  10  and  21  gal.  of washwater had been
 used  for  backwash.  A resume of  the backwash sequences and sand col-
 lection times  for the graded sand  at different expansions is pre-
 sented  in Table 3.

Table 3.  Backwash procedures  for  graded sand.
Wash sequence

Expansion,
%
15
30
40
50
60
75

Porosity in
top 3 in.
0.58
0.67
0.70
0.74
0.77
0.82
Valve
opening,
min
1.0
1.0
1.0
1.0
1.0
1.0
Wash
time,
min
10.5
•6.5
5.0
4.0
3.25
2.75
Valve
closure,
min
2.0
2.0
2.0
2.0
2.0
2.0
Sand collection times

Sample 1,
min
4.5
3.0
2.75
2.5
2.0
2.0

Sample 2,
min
9.0
6.0
5.0
4.5
3.75
3.5
                                 48

-------
The observations made during the course of a run can be grouped into
two categories, (1) data for analysis and (2) data for quality con-
trol.  For series 1 and 2 the following readings were taken as data
for analysis.

During runs A:

a.   The initial effluent quality for each of the three filters at
     intervals of 1 min each for the first 10 rain of a run.

b.   Piezometer readings at depths of 0, 3, 6, and 9 in. for each of
     the three filters at frequencies of one-half hour.

c.   The effluent quality at the total depth of 12 in. for each of
     the three filters at every half hour.  The samples were collected
     at the outlets flowing into the effluent rate controller cham-
     bers.

d.   The effluent quality at intermediate depths of 3, 6, and 9 in.
     for each of the three filters at intervals of every hour begin-
     ning with the first sampling at 0.5 hr after the run began.
     These samples were collected from the continuous drip samplers.

During backwash:

e.   The heights of the expanded bed.

f.   The backwash flow rate.

g.   The piezometer readings at every 3-in. depth of the expanded bed.

h.   The temperature of the washwater.

i.   The backwash water quality from each filter at times of 0.5,
     1.0, 2.0, 3.0, 4.0 and 5.0 min during the washing time of 5 min.

During runs B:

     Similar readings to those taken during runs A, and indicated
     above by a, b, c, and d were taken.

For purposes of maintaining identical conditions from run to run, the
following data were taken for purposes of quality control, during
runs A and B.

j.   Influent iron concentration for one of three filters at fre-
     quencies of one-half hour.

k.   The flow rate through the three filters was monitored and ad-
     justed if necessary every half hour.  Adjustments were only re-
     quired during the latter halves of runs A or B.
                                 49

-------
1.   The room temperature and the water temperature were monitored
     every half hour.  Any small changes of water temperature were
     adjusted.

A similar set of readings were taken during series 3 for the graded
sand, subject to the following modifications.

a.   In order to obtain the peak of the initial effluent quality
     curve, the samples of water were collected at 0.5, 1.0, 1.5,
     2.0, 3.0, 4.0, 5.0, 6.0, 8.0, and 10.0 min, respectively.

b.   Since the depth of sand was 18 in., piezometer readings were
     taken at 0, 3, 6, 9, 12, and 15 in.

c.   The effluent quality was measured at the total depth of 18 in.

d.   The effluent quality at intermediate depths was measured at 3,
     6, and 12 in.

The sampling for backwash water quality from each filter was variable
depending on the duration of the washing sequences as shown in Table
3.  However, the sample collection times were preplanned so that
seven samples were collected from each filter at times corresponding
to usage of equal volumes of washwater.

In order to evaluate the degree of segregation and the distribution
of particle sizes in layers of the fluidized bed, the following ex-
periment was made on the graded sand.  The graded sand bed was fluid-
ized to 507o expansion and allowed to stabilize for nearly one half
hour and then approximately 3-in. layers of the fluidized bed were
siphoned off.  These sections of sand were then oven dried and cooled.
The total dry volumes were measured, and then representative samples
from each layer for purposes of sieving were obtained by repeatedly
reducing the total volume of each layer in a two-way splitting sampler
to about 500 g.  Using the balance sand from each layer after select-
ing the 500 g, bed porosities of each layer were measured by proce-
dures described elsewhere [4].  From the measured volumes and the ex-
panded heights of the same layers, the porosity of each layer when
the total bed expansion was 5070 was also calculated.

For purposes of quality control and to check the general form of the
curves, the following curves were plotted for most of the runs:
(1) head loss against time and (2) ratio of effluent to influent con-
centration against time (i.e., C/CO vs t).  Some typical results are
shown in Figs. 5 and 6.  From the head loss curves it can be seen
that the behavior is reasonably linear for the uniform sand.  The
ratio of effluent to influent curves indicate that even after back-
wash at various expansions the variation from filter to filter is
very small.  Thus, for using the effluent quality as a parameter to
indicate the effectiveness of wash needs a cumulative effluent quality
curve as described by Johnson and Cleasby [67].
                                  50

-------
    8
2   4
             RUN 3B,  TOP 6 in
             Fl  O
             F2 D
             F3 A
                                 TIME,hrs


      Fig. 5.  Head loss curves for run 3B, top 6 in. of filter media.

    Results

    The results  in this  section indicate  the variation of  the  following
    parameters with the  porosity provided during the  controlled backwash
    following the dirtying  run:  (1)  effluent  quality at various depths,
    (2) head  loss increases,  (3) backwash water quality,  (4) backwash
    water volumes, and (5)  sand wash  analysis.

    Cumulative effluent  quality and porosity.   Figure 7 shows  typical
    plots of  the cumulative differential  iron  of the  effluent  with time
    (run 8, series 2).  The ordinate  is the accumulated difference be-
    tween the iron in observation  run B and dirtying  run A for the
    effluent  from the full  depth of the filter.  If the ordinate is nega-
    tive, the cumulative iron in run  A was  greater than that in run B, or
    mathematically  Z   Iron  (B-A)  <  0.  In each run  all three filters
                   time
    were dirtied under identical conditions,  as explained  in the previous
    chapter.   The filters were then backwashed at different expansions,
    and the quality was  studied in the following run.  Study of these
    figures towards the  end of the run will indicate  which filter
                                     51

-------
o
 IU
 tu
    0.06
    0.04
RUN 10B
F10
F2A
F3D
 S  o.oo
 o

 So           1            2            3
                                  TIME, hrs
 Fig.  6.   Variation of  the  ratio  of  effluent  to  influent  iron with  time.
performed the best in a particular run.  The filter with the smallest
positive, or the most negative iron produced the filtrate of the best
quality in the filtration run compared to the dirtying run.

An index called the effluent quality index was defined to compara-
tively grade the filters in each run.  In each run the filter pro-
ducing the best cumulative differential effluent quality was given an
index value of 3, the next best quality was given an index value 2,
and the worst quality was given an index 1.  If two filters performed
almost identically in a run then the two effluent quality index
grades were divided between the two corresponding filters.

The values of the index for all the runs from 7 to 12 and the corre-
sponding porosities in each case are shown in Table 4.

Originally, it was thought that the differential cumulative iron could
be summed from run to run to give the basis for comparison.  In prac-
tice it was found that even with the best possible experimental con-
trols it was impossible to reproduce identical dirtying runs in the
series, due to the inevitable fluctuations in flow rate and iron con-
centrations.  However, within a single run, such changes occurred on
all three filters simultaneously, and comparisons within filters
                                   52

-------
o) -62.5
*»
Z

Z -125
LU
ID
u_
u_
LU
S-187.5
LU
U.
S
A
1A\
\l\
- \
\\
\v
- \V
ti \X.
\v -
NN 0^

RUN 8 POROSITY
0*5*5 P9 /\ — —
0.75 F3 D 	


^oJl^^±^-^---^---A,
Urf
^  -250 —
U
  -312.5
        Fig.  7.   Cumulative differential effluent iron vs time,  run 8.

during a run were the best possible means of study.  Using the above
artificial effluent quality index and repeating the runs so that each
filter was subjected to all the expanded porosities during backwash,
differences due to variations in the sands of the filters were aver-
aged and hence eliminated.  Thus the index provides the best means of
comparison of backwashing effectiveness between different expansions.

The variation of the cumulative effluent quality index (i.e., summing
the index for all runs) with porosity for runs 7 to 12 shown in
Table 4 is plotted graphically in Fig. 8.  The results show a maximum
in the cumulative effluent quality index at a porosity of 0.65 to
0.70.  The maximum indicates that the best effluent quality in the
run following backwash is produced by backwashing the dirty filter at
the expanded porosity of 0.65 to 0.70.

The same type of analysis was done for the effluent qualities mea-
sured at depths of 3, 6, and 9 in. during series 2.  The variations
of the cumulative effluent quality index with porosity for all the
depths of 3, 6, 9, and 12 in. are shown in Fig. 9.
                                  53

-------
 Table 4.   Effluent quality index and porosity for series  2.

Rim
XVU1J
number
7
8
9
10
11
12
Expanded porosity
0.55 0.60 0.65 0.70 0.75 0.78
Backwash ing index
2 3* 1
1 2 3
1 3 2
2 1 3
3 2 l
2 3 i
Cumulative
 effluent
  quality
   index
 The best performance in effluent quality was given in index 3, the
 next best 2, and the worst  1.

The results  for  runs 1 to 6  for the 12-in. depth are presented graph-
ically in Fig. 10.  The results at other depths were not analyzed due
to the fact  that the readings of effluent iron in run 1 were invalid
due to storage of the samples for too long a period before analysis,
as already mentioned under experimental observations.  Since a com-
plete set of readings is required for an unbiased analysis at all
porosities,  the  results for  series 1 could not be analyzed for the
intermediate depths.

The results of the variation of the cumulative effluent quality index
with expansion for the graded sand for the full depth of 18 in. and
for all depths of 3, 6, 12,  and 18 in. are shown in Figs. 11 and 12.
Also marked on figures are the experimentally determined porosities
of the top 3 in. of the graded sand bed while in the fluidized state.

All the results  of series 1, 2, and 3 shown in Figs. 8 to 12 indicate
quite clearly that in every  case the best effluent quality in the run
following backwash is obtained by expanding the bed to porosities of
0.65 to 0.70 during backwashing.

Initial effluent quality and porosity.  A technique similar to above
was used to study the variation of initial effluent quality with po-
rosity during the preceding backwash.  The cumulative initial efflu-
ent index was given a value of 3, 2, or 1, depending on the cumula-
tive differential iron between runs B and A during the first 10 min
of a run.  The method was identical to that used to evaluate the
                                  54

-------
     8
t
g
o
»-
LU
u
          RUNS 7 THROUGH 12
          UNIFORM SAND
          12-in.  DEPTH
                     I
              I
I
I
        0.55
0.60        0.65        0.70        0.75
 AVERAGE POROSITY DURING BACKWASH
                    0.80
 Fig. 8.  Cumulative effluent  quality index vs porosity, series 2, 12-in.
          depth.

 effluent quality as already described.  The results indicate no re-
 lationship between the  initial effluent quality in the run following
 backwash and the porosity  of  the expanded bed during backwashing.
 It was felt that the initial  effluent quality was not dependent on
 the backwash expansion  but was a function of the rate of closure of
 the backwash valve.

 Head loss increases and porosity.  A study was also made on the
 effect of backwash on the  head losses in the run following backwash.
 Again comparison was made  between  filters based on the difference of
 head loss in run B over that  of the dirtying run A.  The results were
 not conclusive and are  not presented here.

 Backwash water quality  and porosity.  The parameter which provided
 data that was the most  consistent  in all the runs was the backwash
 water quality.  It enabled comparisons between different backwash
 porosities to be made on the  basis of usage of equal quantities of
 washwater, even though  the actual  washwater used in the series 1 and
 2 was dependent on the  constant wash duration of 5 min.  For series 3
                                   55

-------
      28
      26
      24
>;

<

O


UJ
Z>
      22
      20
   D
   U
              SERIES 2
              RUNS 7 THROUGH 12
              UNIFORM SAND
              ALL DEPTHS
                                               SERIES 1
                                               RUNS!THROUGH 6
                                               UNIFORM SAND
                                               12-in. DEPTH
         0.55        0.60       0.65        0.70       0.75
                      AVERAGE POROSITY DURING BACKWASH
                                                          0.80
Fig. 9.  Cumulative effluent quality  index vs porosity, series 2,
         all depths,   (above)

Fig. 10.  Cumulative effluent quality index vs porosity, series 1,
          12-in. depth,  (below)
                                 56

-------
2
2
    8
I
C

UJ
2  4
RUNS 20 THROUGH 25
GRADED SAND
18-in.  DEPTH
                                 I
               20       30       40       50
                           PERCENT  EXPANSION
              J	I	I	I
                                         60
 70
80
         0.58
             0.67      0.70     0.74      0.77

                 POROSITY OF TOP 3 in.
0.80   0.83
 Fig. 11.  Cumulative effluent quality index vs  expansion,  series 3,
           18-in. depth.

 the total volume of washwater used  for the different expansions was
 maintained the same for  all the runs  by varying the durations of
 wash.

 Figures 13 and 14 illustrate the backwash water quality  for  series 1
 in terms of the iron concentration  in mg/1 in samples  of washwater as
 a function of the total  volume of washwater used up to the time of
 sample collection.  Using the time  of collection of samples  and the
 flow rate during that particular wash the total washwater  used was
 calculated and plotted as the abscissae.  The plotted  points are from
 different filters and different runs  but  are  grouped together to in-
 dicate the variation of  backwash water quality  with porosity.  The
 apparent scatter in the  points towards the end  of the  backwash is due
 to the graphs being plotted on logarithmic coordinates.  The loga-
 rithmic coordinates were necessary  to show the  variations  in backwash
 water quality which range from 1000 to 0.2 mg/1.  However, for pur-
 poses of analysis the most relevant sections  of these  primary curves
 shown in Figs. 13 and 14 are the lower curved portions before the
 curves reach asymptotic  values.  Magnified curves of these sections
                                   57

-------
    31
    27
O
t—
LU
23
r  19
<
U
    15
           RUNS 20 THRO UGH 25
           GRADED SAND
           ALL  DEPTHS
                                 1
                                      1
               20
                   30       40       50
                       PERCENT EXPANSION
                                 1
                                     1
60
1
70
80
          0.58
                  0.67      0.70      0.74      0.77
                      POROSITY OF TOP 3  in.
        0.80   0.83
Fig. 12.
      Cumulative effluent quality index vs expansion, series 3,
      all depths.
for series 1 are shown in Fig.  15 in arithmetic  coordinates.  The
lines drawn are smoothed curves through the  means  of  the values from
the three filters.  The curves  represent the mean  variations of back-
wash water quality with volume  of washwater  at the different poros-
ities.

The smoothed curves of Fig.  15  were used to  prepare secondary curves
showing the variations of final backwash water quality with porosity
for constant volumes of total washwater (Fig. 16).  The points
plotted are the intersections of ordinates at washwater volumes of 20
and 25 gal., respectively, with the smoothed curves drawn in Fig. 15.
                                  58

-------
    100

     60
o    30
ce.
     10
§  6.0
UJ
    3.0
 <
    0.6
    0.3
     0.
                                        POROSITY  RUN
                                            0.55
                   0.65
       3  Fl  ®
       2  F2 •
       4  F30
       2  Fl A
	5  F2A
       6  F3 A
0.75   4  Fl B
	6  F2 •
       2  F3 D
1
                         10         20          30
                             WASHWATER VOLUME, gal.
                                  40
      Fig. 13.  Backwash water quality vs washwater volume,  series  1.
                                   59

-------
    200


    TOO

    60
z
o
>:    io
g  6.0
LLJ

S


X
   3.0
   0.6
   0.3
   0.1
                                            POROSITY RUN
                                              0.60
       1
       4
       5
                                              0.70


                                              0.78
	1  F2
       3  F3
       5  Fl
    - 3  F2
       1  F3
OMIT
                                                         F2
                                                         F3
       6  Fl  OMIT
                                                             A
                                                             S
                        I
 I
                       10          20          30

                       WASHWATER VOLUME, gal.
                                                          60
    Fig. 14.  Backwash water quality vs washwater  volume,  series 1,
                                  60

-------

3


Z 2
9

^
> i
t
_i
<
z>
Q£
LLJ
1 °

i
<^^ O
3
g
0
CD
<
2 2
^^
S
i—
1
0
1 \ I
1 \ U POROSITY RUN
l \ r^
— \ \ i \ 0.60 1
1 i \ 1
\ \ \ »
\ \ \ 0.70 6
% * I
- \ \ \ — 1
\ \ \ 0.78 5
\ \ \ — — — 3
x^ \ \ '
S^ ^N A
B ^^-\
•^ "^m
**• ^NSJ
* ^33^.^^ 	

i i ^""^
i ' \
1 » \
i i \
> 1 \ 0.55 3
— i \
\ » \ 	 4
^ \ \ 0.65 2
\V\ —2
\ \ 0.75 4
\ \ 	 6
\\
\ \A
\\
V^
^^i,
^•^-^r
i i i"
10 20 30
F1 OMIT
F2 •
F3 0
Fl OMIT
F2 A
F3 A
Fl B
F2 •
F3 D



f51

' 1



Fl ®

F2 •
F3 0
Fl A
F2 A
F3 A
FT B
F2 •
F3 D

1
40
                      TOTAL WASHWATER,  gal.
Fig. 15.   Backwash water quality vs washwater volume, series  1.
                              61

-------
   2.4
Z

S£  2.0
   1.6
O
O£
LU

I
I
<
   1.2
   0.8
   0.4
   0.0
I
I
                 TOTAL WASHWATER USAGE
                    • 25 gal.
                    A 20 gal.
I
          0.55       0.60        0.65       0.70        0.75
                   AVERAGE POROSITY DURING  BACKWASH
  Fig. 16.   Terminal backwash water quality vs porosity, series 1,
                                 62

-------
The results show once again that the best terminal backwash water
quality is achieved by backwash at the porosity of 0.70.  This has
resulted from analyses which consider backwashing to different ex-
pansions, but using a constant total volume of washwater.  The re-
sults indicate that most effective backwash is achieved by expansion
to porosities around 0.70.

An alternative graph, also derived from Fig. 15 by considering the
different volumes of washwater needed at the different expanded po-
rosities to achieve a given terminal backwash water quality, is shown
in Fig. 17.  A family of curves for terminal backwash water qualities
of 0.75, 1.00, and 1.50 mg/1 of iron is shown.  These graphs show
again that the minimum quantity of washwater necessary to obtain a
given terminal backwash water quality occurs at the porosity of 0.70.

The above analyses should be restricted to the lower sections of the
curves when the quality changes become small, since only in these
sections are the results meaningful.  Identically, similar graphs
resulted in all the experiments of series 2 for the uniform sand and
of series 3 for the graded sand.  In every graph a minimum in the
total washwater volume usage or the terminal washwater quality oc-
curred around the anticipated porosity range of 0.65 to 0.70.  Thus
both of these parameters have provided still further evidence of the
optimum theory developed in a previous chapter.

Physical sandwash and porosity.  As already recorded, an extra param-
eter to evaluate the effectiveness of backwash was proposed based on
the amount of iron removable from the sand by a physical wash.  The
washing procedure was simple abrasion using a magnetic stirrer under
standard conditions.  Cpnsiderable amounts of iron were removable
from the sand by this method, providing final evidence for the fact
that negligible collisions and abrasions between particles occur in a
fluidized bed.  If there were considerable abrasion in the fluidized
state, it should not be possible to remove these large amounts of
iron by a physical wash.

The iron removable from the graded sand in mg/g as a function of ex-
pansion is shown in Fig. 18.  The points plotted are for the first
two runs on the graded sand - runs 20 and 21.  These runs were the
initial runs made on the new graded sand after it had been subjected
to one unnumbered run for purposes of coating the new sand with at
least a small layer of the iron floe.  Though similar measurements
were made for all the runs of series 3, it was found that the results
of runs 22 to 25 were subject to considerable error due to the fol-
lowing cause.  In series 3, the influent suspension contained 7 mg/1
iron and 0.10 mg/1 of a nonionic polyelectrolyte.  As series 3 pro-
gressed from run to run, mudballs started building up due to the
added polyelectrolyte.  These were about 0.5 to 2.0 mm in size and
consisted entirely of globules of the precipitated iron without any
sand within them.  They floated on top of the sand layer during flu-
idization, and every time the sample of sand was drawn for analysis
                                  63

-------
 o
 o>
o
85
I
to
I
    26
    24
22
20
    10
    16
                       I
                                TERMINAL WASHWATER QUALITY
                                      0.75 mg/l IRON  •
                                      1.00 mg/l IRON  A
                                      1.50 mg/l IRON  •
                                         I
           0.55        0.60        0.65        0.70       0.75
                  AVERAGE POROSITY DURING BACKWASH
      Fig.  17.  Backwash water volume vs porosity, series  1.
                                64

-------
a
§
oe
o
z
o
c*
   0.10
   0.09
   0.08
   0.06
   0.05
   0.04
              0.58
                                                     • RUN 21
                                                     O RUN 20
10      20       30      40       50      60
                    PERCENT EXPANSION
 I	  I         I        I         I        I
                0.67     0.70     0.74     0.77

                   POROSITY OF TOP 3 in.
                                                              70
0.80
Fig. 18.  Iron removable by physical abrasion test vs  expansion,  runs
          20 and 21.
                                   65

-------
 considerable amounts of these mud balls or globules  were  drawn with
 the sand sample and caused the iron removable readings  to be  erratic
 from run to run.   We found that as the series of runs progressed  the
 mud balls became  bigger and bigger, causing larger and  larger errors.
 Even though a modified  larger expansion wash, nearly 100%, was used
 for a few minutes during the standard  cleaning at the start of each
 run, removing substantial amounts of the mud balls was  still  not
 possible.

 For the  above reason the physical abrasion test  results of runs 22
 to  25 were invalid and  are not presented.   However,  in  runs 20 and  21
 the mud  balls were still few in number and small in  size  and  did  not
 substantially affect the readings.

 The results shown in Fig.  18 reconfirm the fact  indicated in  the
 other sections that an  expansion of about  40 to  50%  produces  the
 cleanest sand in  the graded filter.  These expansions cause porosi-
 ties of  about 0.70 to 0.74 in the top  3-in.  layer of the  expanded bed
 of  graded sand.

 Analysis.   The results  presented in the  preceding pages prove beyond
 little doubt  that an optimum backwash  occurs in  a system  expanded to
 achieve  a porosity of approximately 0.65 to  0.70 in  the layers con-
 taining  the most  amounts of suspended  matter.  In the case of a uni-
 form sand bed an  expansion to a porosity of  0.70 is  equivalent to
 nearly 90% expansion of its  height.  For a theoretical uniform sand
 consisting of identical particle  sizes  this  expansion is  an exact
 and  fixed condition.  However,  in filtration practice uniform sand
 beds are  never used,  nor is  it  feasible  to provide for 100% expan-
 sions  because of  the  unreasonably large  backwash flows required.  We
 are  fortunate that  both these  limitations  are  simultaneously  removed
 by use of a graded  sand.   A  40% expansion  of a typical graded filter
 sand can  be shown to  cause the  porosity  of the top 3-in.  layer to
 reach  a value of  0.70,  and use  of three  different parameters  has
 shown  that the optimum  backwash  for  the  graded system occurs  at
 expansions  of about 40  to  45%.  These values  of  expansion can be  ob-
 tained in  practice.

A few points  need  to be  made  regarding this  optimum  of 40 to  50%  for
 graded systems and  the  results  reported  in the sanitary engineering
 literature.   The  fact that effective backwash  requires an expansion
 of about  50%  for  graded  systems has been a well  known rule of  thumb
 in filtration design  and practice.  This rule developed originally
 from the work of Hulbert and Herring [62]  in  1929.

 Several workers have  suggested during the  last decade that expansions
 of 20  to 25% may be sufficient  for effective backwash [10,26,67,132].
None of these papers, however, provide fundamental considerations or
experimental  results which are valid to draw this conclusion.   Baylis
 [10] suggested the figure without any experimental work.  Camp et al.
 [26] reported this expansion  as suitable for all filters on the basis
                                  66

-------
that serious problems did not occur in the operation of the Billerica
water treatment plant.  But note that the Billerica plant had multi-
media filters, and expansions of 20 to 25% easily give porosities of
about 0.70 in the top coal layer.  This can be seen in the results
reported by these workers themselves.  This, in fact, is additional
evidence for the hypothesis of this report.

Thus, careful analysis indicates that the results reported in the
literature [26,62] are consistent with the theory and experimental
work of this report.  Also, remember that effective backwashing does
not necessarily mean optimum backwashing, and due to the rather flat
nature of the shear stress maximum it is possible to backwash filters
effectively even though the optimum condition is not obtained.  In
case one may be tempted to run away with the idea that a lower ex-
pansion may result in a saving of washwater, Fig. 17 needs to be re-
membered.  It clearly shows that backwashing at lower expansions than
the optimum necessarily results in the usage of larger amounts of
washwater to achieve a given bed cleanliness.

A concluding summary.  The results summarized and the analyses pre-
sented in this section give a complete picture of optimum backwashing
by water fluidization alone.  The experimental results are entirely
consistent with the theory of optimum backwash developed in a pre-
vious section and provide excellent confirmation of the theoretical
results.  Optimum backwash has been shown to simultaneously provide
these advantages:  (1) a better effluent in the following run, (2) a
minimum usage of washwater, and (3) a minimum growth of the coatings
on the sand.  This plurality of advantages should considerably improve
the performance of most filtration plants, if optimum backwashing
is put into operation.
                                  67

-------
   VII.  WASTEWATER FILTRATION AND  BACKWASHING - LITERATURE REVIEW
Filtration has been used  in  the United States as  a  liquid  and  solids
separation process for wastewater  since  1883, but was not  widely im-
plemented over the years  because settling  and biological treatment
processes were considered adequate for the needs  of the times.  How-
ever, with stringent  federal and state effluent standards  presently
reflecting the demands of the citizenry  for more  complete  treatment
of wastes, filtration is  becoming  increasingly popular.  Similar re-
quirements for better waste  treatment were necessary in Britain nearly
25 years ago, so  the  majority of what has been accomplished in waste-
water filtration  and  backwashing progress  since 1949 has been  the
result of British research,  development, and experience.

Experience with filters used in water treatment has demonstrated the
need for effective media  cleaning  techniques.  A  similar need  exists
with respect to wastewater filtration, but the problem is  greatly
compounded because of the variable characteristics  of sewage.  Often
sticky and gelatinous, the solids  removed by the  media are much more
resistant to cleaning procedures than those normally encountered in
water treatment.

The following literature  review summarizing wastewater filtration
and backwashing experience is a summary of a more comprehensive re-
view prepared by  Rice [99].

A diverse array of designs is available for wastewater filtration,
varying in flow configuration, bed depth, media type and gradation,
and performance.  Cleaning methods, however, generally rely on the
application of water  or air  and water to remove entrapped  solids from
the bed (chemicals such as chlorine occasionally  are used  as cleaning
aids).  The purpose of this  section of the review is to review in a
general way the wastewater filter  designs and cleaning techniques.

          Types of Wastewater Filters and Cleaning  Techniques

Conventional Rapid Sand Filtration

Because of their use  for  years in  filtering water for potable use, it
is not surprising that rapid sand  filters were among the first to be
used in wastewater filtration.  Bed depths of 6 to  36 in. have been
reported [13,41,87,121,136]  for full-scale applications, but current
practice favors depths of 24 in. or greater.  The gradation of sand
used in the installations varies widely with location, but research
and operating experience have demonstrated that the  relatively fine
sand, approximately 0.5 mm, used in water treatment  is unsuitable for
wastewater filtration because of rapid head loss buildup and conse-
quently shorter filter runs.  Considerable pilot  scale research has
shown that 1 to 2-mm media size will produce good effluent quality
and allow reasonable  filter runs [59,68,129].
                                  68

-------
Design of these facilities closely parallels that employed for years
in waterworks, the filters placed in rectangular concrete boxes
equipped with perforated block, nozzle type, or header and lateral
underdrain systems [41,87,136].

Backwash water is usually drawn from a filtered wastewater storage
tank although unfiltered water is occasionally used.  In either case
the washwater is pumped through the bed in an upward direction, carry-
ing the accumulated solids upward to washwater collection troughs.
Washwater is ordinarily applied at a rate sufficient to expand the
sand bed 10 to 20% during fluidization, usually at rates from 15 to
25 gpm/sq ft.

In Britain, air is almost universally used in wastewater filtration
plants as a media scouring aid prior to or in the course of intro-
ducing the washwater.  The air agitates the media and helps to break
up agglomerations within the bed, thus allowing the water backwash to
more easily remove the entrapped solids.  Although installed less
frequently, rotary surface washers have been used for the similar
function [87,88].  No reports of backwashing studies testing the
effectiveness of either of these two scouring techniques were dis-
covered, and  since wastewater filtration plants in the United States
are rare, little direct evidence exists as to which method best
cleans the media.  Indications from scattered passages in filter per-
formance reports are that omission of either air scour or rotary sur-
face wash has led to difficulties [59,68,118].

Dual- and Triple-Media Filters

Dual- and triple-media filters are being used to provide filtration
from coarse to  fine media size in the direction of flow, and thus to
achieve longer  filter cycles without detriment to filtrate quality
[38,59].

The most common type of dual-media filter is an anthracite and  sand
design although other combinations such as activated carbon and  sand,
resin beds  and  sand, and resin beds and anthracite have been re-
ported  [82].  Studies [59,127,128] appear to conflict somewhat  on the
degree to which solids penetration and efficient bed utilization
occur in a  dual-media filter.  Bed depth, flow rate, media sizing,
and  the nature  of  the filter  influent  (activated sludge versus  trick-
ling filter effluent) appear  to  be variables influencing both  the
penetration and degree of removal in  a dual-media filter  [38].

As with  any granular media  filter, efficient backwashing  is essential
 to prevent  deterioration of the  filter bed  condition, which results
 in filter  cracks,  mud balls,  high initial head losses and reduced
 filter  runs or  poorer  filtrate quality.  Backwashing  techniques com-
monly used  for  dual-media  filters are essentially the same  as  those
 described  for rapid  sand  filters.  Air scour prior  to backwashing is
 common  practice and  is  introduced through  the  nozzle underdrain
                                   69

-------
 systems common with this type of filter [58] .   The use  of a series  of
 air-scour and rest cycles,  referred to as  pulsed air scour,  has  been
 recommended as a means  of improving the efficiency of media agitation
 [58,86].   Initially the air will follow the  path of least resistance
 and may completely skirt agglomerations in the  bed if continued.  By
 pulsing the air and allowing the media to  resettle,  areas of the bed
 resistant to break up are continually  lifted and dropped,  resulting
 in better separation of the entrapped  solids from the filter media.

 Triple-media filters are the result of a logical extension of the
 principles outlined for dual-media  filters.  The most frequently used
 design  incorporates anthracite,  sand,  and  garnet,  which are  sub-
 stances having approximate  specific gravities of 1.7, 2.65,  and  4.2,
 respectively.   Gradation from coarse to fine follows, obviously, the
 same sequence,  and it is claimed that  the  filter is  less  susceptible
 to shock  from rapid fluctuations in suspended solids  concentration
 [111].  It has also been suggested  that triple-media filters are
 superior  to deep-bed filters using  a single coarse media  [111],  but
 reliable  pilot-scale studies have not  demonstrated superiority in
 effluent  quality or process reliability [68].

 Promoters of the triple-media filter have  expressed  concern  about
 the use of air scour in this and other filters  [111].   Disadvantages
 of air  scour are listed as  increased downtime,  possible media loss  or
 bed upset,  and  complication of the  backwash cycle.  The use  of
 rotary  surface  washers  is recommended  as an alternative.   Others have
 suggested that  air scour is  perhaps  the only practical  method  for
 cleaning  deep bed  filters like the  dual- and triple-media units
 [38,86],  and  that  increased  scouring efficiency is possible  with air.
 Filtered  water  is,  in any case,  generally  used  as  washwater  and  is
 applied at  a rate  of about  15  gpm/sq ft.

 The  use of  triple-media filters  for waste  treatment  is  growing in
 popularity  in  the  United States,  with  over 50 installations  in opera-
 tion to date.  Most  designs  have  favored bed depths of  36  to 42  in.
 and  hydraulic  loading rates  of 5  to  6  gpm/sq ft.   Pressure filters
 are  normally used  for treating secondary effluent  from  plants  with
 flows of  less  than 5 mgd  and gravity filters for  treatment works with
 flows in  excess  of  this  figure [43].

 Immedium  Upflow Filter

Although  relatively unknown  in the United  States,  the Immedium filter
was  invented in  1961 by  a Dutchman named Pieter  Smit  and was  subse-
quently developed by the Boby Corporation  in Britain  [9,13],   It is
 a deep-bed, high rate,  upflow filter using a patented grid placed
 several inches below the top of  the bed to prevent the  expansion of
 the media during filtration.  Typical  arrangements for both  pilot-
 and  full-scale  Immedium  filters  are presented in Figs.  19  and  20.
The  grid  is the key  to  the  success of  the design and consists  of a
 series of parallel bars  at 4  to  6-in.  centers which normally provides
                                  70

-------
ACTIVATED
SLUDGE
WORKS FINAL
EFFLUENT
      WET WELL
        PUMP
          FLOW
          METER
      MANOM-
        ETER
OVERFLOW
                              EFFLUENT
                                WEIR
                              ft  COARSE
                           &SAND (1-2 mm)$
                           •COARSER GRAVEJ.
                           UNDERDRAIN-
                                           BACKWASH
                                           OUTLET
                       DRAIN
                                             MEDIA
                                            SURFACE
SAMP
TA
Lll^
NK
JG
                                                           TO
                                                           WASTE
                                         AIR BLEED
       Fig. 19.
    Pilot-scale Immedium filter used at West
    Hertfordshire, England  [142].
                               71

-------
CONCRETE
 FILTERED
 WATER

   WASTE
          TnTTTTmiLmTTTTTTTmrrniiirimniniiiiTTiin
'SAND RETAINING GRID


NOZZLE DISTRIBUTION SYSTEM
                                                       AIR
                                                          -XJ—
                                                          INLET
                                                          -txj-
                                                          WASH
                             OPEN TYPE
SAND RETAINING
   STEEL SHELL
             \
                   i	n
                         rfv.
                                                    FILTERED
                                                    WATER
                                      WASTE
                                N
     INLET
       WASH
                                   NOZZLE DISTRIBUTION SYSTEM
    Fig.  20.  Immedtum filter arrangements for full-scale
             installations, open and pressure [l3].
                               72

-------
openings 100 to 150 times the size of the smallest particle contained
in the bed [13,58].  During the course of filtration the sand forms
compression arches in the vicinity of the grid which resist the tend-
ency to lift or fluidize the bed, even at rates up to 6 gpm/sq ft and
at pressure drops equivalent to 20 ft of water.

Intense interest in the Immedium upflow filter in Great Britain has
resulted in at least five independent studies [68,83,86,131,142],
which are discussed in more detail later.  Typical media gradations
and depths, in the direction of flow through the filter, are 4 in. of
10 to 15-mm gravel, 10 in. of 2 to 3-ram gravel and 60 in. of 1 to
2-mm sand, and the filter can be operated in a pressure or open type
housing as shown in Fig. 20 [9,13,58].

Backwashing is usually initiated when the head loss reaches 6 ft and
is preceded by draining the water above the filter to a level just
above the media surface and air scouring the filter at a rate of
5 cfm/sq ft.  The air serves also to break up the sand compression
arches at the grid, and the backwash water is then turned on to ex-
pand the bed before the arches can reform.  Since backwash water is
applied at 16 gpm/sq ft in the same direction as the influent, the
latter is conveniently used as washwater in most instances.  Expan-
sionsof 27o [9] to 20% [13] have been reported as common during back-
washing, but it is the writer's opinion that the former value is more
accurate for 1.0 to 2.0-mm sand fluidized at 16 gpm/sq ft.

The first reported full-scale Immedium filter plant became operation-
al in 1969 at the East Hyde Works in Luton, England following a
series of pilot-scale studies.  As a result of the pilot-scale stud-
ies, the backwashing procedure was modified to provide a series of
alternating high and low rate air scours and washwater applications,
both of which have been automated [9,82].

Deep-Bed, Coarse Sand Filters

The use of deep-bed filters with a single media of coarse sand has
been developed in Germany.  According to Jung and Savage [70] its use
is widespread in potable water treatment, with over 200 existing in-
stallations.  They are also being promoted for wastewater treatment
with media depths of 4 to 6 ft and media sizes of 1 to 2, 2 to 3, and
3 to 6 mm for wastewater filtration  [108],  Sand is very uniform,
with a uniformity coefficient of 1.25 or less.  Backwash is first,
with air  and water simultaneously followed by water alone.  The 'back-
wash rate is 6 gpm/sq ft during the  simultaneous air-water phase  and
8 gpm/sq  ft during the water only phase.  Air-scour rates are 6  scfm/
sq ft for the 2 to 3-mm and 3 to 6-mm sand and 3 scfm/sq ft for  the
1 to 2-mm sand.  Adequate freeboard  to overflow level  (20 to 30%  of
bed depth) is required to prevent media  loss with the  finer sand.
Backwash  is for extended periods of  15 to 20 min so that total wash-
water used per backwash is not reduced compared to higher rates used
in United States dual-media filters.
                                   73

-------
The backwash described above is well below fluidization velocity for
the media of the sizes indicated.  Thus, the mechanism of cleaning
comes  into question.  Jung and Savage  [70] describe the air and water
delivery systems and the presumed action in some detail.  Air is dis-
tributed evenly by a pipe network.  Water is delivered by a no (very
low) head loss precast block underdrain floor.  The rising columns of
air act as airlift pumps to ensure uniform water distribution.  The
localization of the air columns causes the velocities of air and
water  flow to vary from zero to quadruple the mean and results in
vigorous pulsating washing action.

Quantitative studies demonstrating the effectiveness of this back-
washing system were not found.

Automatic Backwash Filters

In order to backwash any of the wastewater filtration systems pre-
viously discussed, it is required that they first be removed from
service.  However, the automatic backwash (ABW) filter, as manufac-
tured by the Environmental Elements Corp. (formerly the Hardinge
Division) of Koppers Company, Inc., allows both filtration and back-
washing to occur in the same bed.  A full-scale version of this sys-
tem, similar to that shown in Fig. 21, is now in operation at
Chicago, Illinois, and is filtering chemically treated secondary
effluent in two 12-in. deep sand beds having a total surface area of
1329 sq ft.  The filter beds consist of a series of 8-in. wide, con-
tiguous compartments separated by steel plates which run perpendi-
cular to the long axis of the filter.  A carriage assembly traveling
on rails mounted on the filter box walls suspends a cleaning hood
above the compartmented bed for the full width of the filter [116].

Backwashing is accomplished by centering the cleaning hood over a
given compartment, thus isolating it and allowing the rest of the bed
to continue filtration.  A port on the side of the isolated compart-
ment is opened to the moving backwash supply pump and shoe, and the
washwater is pumped into the compartment underdrain and up through
the sand.  Another pump mounted on the cleaning hood withdraws the
spent washwater and discharges it to waste.  The time required to
completely wash all the compartments in one of the 53.2 by 12.5 ft
beds at the Chicago installations is 57 min [116].

The principal investigator recently (1975) visited this Chicago in-
stallation at"Hanover Park after it had been in service about seven
years.  During that period the filters had been used for direct fil-
tration of a prechlorinated secondary effluent from an activated
sludge plant which was sometimes coagulated prior to filtration.  The
plant superintendent, Mr.  Robert A. Ziols reported satisfaction with
the filters and no particular problems related to inadequate back-
wash.  The automatic backwash occurs about nine times per day.  They
have routinely replaced about 1 in. of filter sand each year.  In-
spection of the media at such times had not revealed typical dirty
                                  74

-------
      CARRIAGE DRIVE
             MOTOR tk
         WASHWATER
          PUMP
        BACKWASH
        PUMP
                          WASHWATER
                            LAUNDER
                                           UQDTD
                                           CONTROL
                                         LEANER
WATER LEVEL
       •EFFLUENT CHANNEL              INFLUENT CHANNEL

      Fig.  21.   Environmental Elements Corp.  (Koppers) full-
                scale automatic backwash filter (from
                manufacturer's brochure).
filter problems such as mud balls and filter cracks.  An expansion of
the plant will include additional filters of the same type.   The
manufacturer now uses sand with an effective size of 0.6 to  0.65 mm
for wastewater filters and reports 25 wastewater installations be-
tween 1966 and 1974.

Moving-Bed. Continuously  Cleaned Filters

In the next few years wastewater filtration practice may encompass
the use. of moving-bed filters which are at present in the research
and development phase.  One filter of this type is basically a verti-
cal,- upflow unit in which provisions have been made to.continuously
remove a small portion of the bottom-most sand, the media making
first contact with the influent.  The sand is then cleaned and lifted
to the top of the'filter  where it is redeposited.  In this manner
the flow through the filter contacts progressively cleaner media as
it travels through the bed, and the dirtiest portion of the  media is
constantly being drawn off  and cleaned.  Preliminary results indicate
that raw water quality has  little effect upon performance for flow
rates comparable to conventional upflow and downflow units [58],
                                 75

-------
 A variation of the moving-bed filter known as the  Simater  filter  is
 being manufactured by Simonacco Limited of Carlisle, England,  and was
 tested in a pilot-scale study in Derby, England  [68],  Figure  22
 shows the basic arrangement  of the  unit, which differs from the pre-
 viously described  filter in  that the unit  distributes the  influent
 radially from the  center of  the bed (A and B).   Vertical perforated
 pipe  laterals whose orifices are covered with filter cloth are spaced
 around the outer circumference of the filter  housing to collect the
 effluent, which is then discharged  through lines D  and E.   Dirty
 media is withdrawn in line F and air lifted to the  top of  the  filter
 where it is discharged into  the upper chamber, G.   During  the  lifting
 processes the solids are removed from the  media  by  the scrubbing
 action so that the sand falling to  the top of the  filter has been
 cleaned.  The solids are then withdrawn from  the upper chamber at
 point G [68].

                             Case Histories

 The following case histories will present  the backwashing  procedures
 and their effectiveness as reported in various wastewater  filtration
 studies from the literature.   In many of these papers, substantial
 filter performance data are  also presented but will not be repeated
 in this backwashing report.   The performance  data have also been  sum-
 marized by Rice [99].

 Early Practice According to  Streander

 Perhaps one of the unsung pioneers  of wastewater filtration was
 Philip B.  Streander,  a New York City consulting  engineer whose in-
 terest in sewage filtration  began in the late  1920's.  In  a solitary
 article published  in 1935 [121]  and a subsequent series of three  ar-
 ticles published in 1940, Streander [122,123,124] outlined the state-
 of-the-art of sewage  filtration in  this country  and abroad.

 Streander proposed a mechanically cleaned  downflow  filter  [121] of
 silica sand  or crushed  anthracite graded to 0.59 to 0.84-mm particle
 size.  Filtration  rates  of 2  gpm/sq ft were recommended, and cleaning
was accomplished by a moving,  full-width cleaning head.  The head was
 equipped with  two  rows  of hollow rake  teeth, positioned to break up
 the surface  and  subsurface portions  of a 6  to 18-in. deep  filter bed.
The teeth  were equipped with  orifices  through which high pressure
washwater  was  pumped,  causing an expansion  of the media directly
under  the  cleaning heads  equivalent  to a rise rate  of 16 in./min.
 Solids  entrapped by the media were  released and  carried above  the ex-
panded  bed to  an upper  zone where they were withdrawn in the spent
washwater  by  a pump mounted  on  top  of  the hood.

Although  it was thought  that  the  combined mechanical and hydraulic
 action would be sufficient to  thoroughly clean the  filter,  perfor-
mance  in  actual installations proved  otherwise.  It was discovered
that  the  cleaning method was highly  efficient in the upper part of
                                  76

-------
                 SEE TEXT FOR EXPLANATION
             BAFFLE
      H
  SURFACE AT MEDIA -


           HOUSING.
   INFLUENT
      A
                          *^*~*~i*~i~*~*j(
                           WATER LEVEL
                          CENTRAL INFLUENT
                            DISTRIBUTOR
                                             MEDIA
Fig.  22.
                                                       AIR LIFT
Pilot-scale "Simater" radial-flow, moving-bed  sand
filter by  Simonacco Ltd.  of  Carlisle, England  [68].
                              77

-------
 the bed but  could  not effectively clean the  media directly  above  the
 supporting screen.  As a result  bacterial  slimes  developed  which  ad-
 versely affected the  capacity and efficiency of the  filter  [122],

 Much of Streander's thinking  was directed  along the  lines of  shallow-
 bed silica sand filters employing mechanical cleaning devices.  He
 felt these designs had definite  merit  as a secondary treatment  system
 for the removal of solids escaping the primary settling process.
 Because of the automatic backwash feature, installations of this
 equipment  for direct  filtration  of primary effluent  had consistently
 removed 50%  of the incoming suspended  solids regardless of  solids
 concentration or filter hydraulic loading  (apparently 2 gpm/sq  ft was
 maximum)  [124].  Streander recommended bed depths of 6 to 12  in.  and
 hydraulic  loadings less than  or  equal  to 2 gpm/sq ft.

 Although a proponent  of shallow-bed, mechanically cleaned filters,
 Streander  was also aware of the  potential  of backwash type, deep
 granular filters in sewage treatment.   He  reported on the full-scale
 installation of such  a plant  in  Wuppertal, Germany,  a city  of about
 410,000 people at  that time [122],  As a result of extensive  pilot-
 scale studies of filtration rate,  sand size,  and  bed depth, the de-
 signers of the Wuppertal plant chose a 1.0 to 2.0 mm sand placed  to
 a depth of 28 in. over a 4-in. layer of coarse and fine gravel.  Ten
 filter beds  were provided, each  26  ft  3 in.  by 123 ft long, giving a
 combined filtering area of 32,280 sq ft for  a flow rate of  24 mgd.
 At  the relatively low hydraulic  loading rate of 0.5  gpm/sq  ft and
 filtering  primary effluent, suspended  solids  removals averaged  only
 40£.   Streander felt  that the low efficiency was  the direct result
 of  a poor  backwashing procedure  which  did  not thoroughly clean  the
 filter.  Basically the backwash  sequence consisted of air scour of
 unspecified  rate and  duration followed by  a  low rate backwash of 6 to
 8 in./min  (vertical superficial  velocity).   Streander speculated that
 the  air agitation loosened the heavier trapped solids, but  the  low
 wash rate  could not effectively  carry  these  away  nor remove the
 finer,  more  tenaciously bound solids.

 Regardless of the depth of bed employed, Streander realized the need
 for  proper sand size  selection and  the dependence of such upon the
 nature  of  the influent solids, the  concentration  of  the solids, and
 the  availability of washwater [122].   He stressed that media  selec-
 tion  should be carefully considered to prevent premature clogging and
 to promote distribution of the suspended solids removal throughout
 the  filter bed.  Furthermore, he  recognized  the necessity of  adequate
media  cleaning techniques with respect  to proper  filter operation.

 Streander predicted the  increased use  of filtration  as an effluent
 polishing  application for  treating  activated  sludge  and trickling
 filter plant effluents.  Placed  after  final  settling tanks, filters
 could, he believed, be used to remove  the  lighter, less settleable
 solids, thus permitting  a  smaller clarifier or overloading of an
                                  78

-------
existing unit.  The following quote [124] summarizes his views on the
potential of wastewater filtration:

     It is the writer's belief that this matter of dollars and
     cents value of effluent strainers has only begun to be
     realized, and that the future will find these worthy ad-
     juncts to treatment plants coming into widespread use as
     an economy feature as well as being producers of more
     even and dependable quality of plant effluent.

British Studies

Luton.  British wastewater treatment literature abounds with refer-
ences to the Luton Corporation Works filtration plant and to experi-
mental filtration studies conducted there.  In 1949, prior to the in-
stallation of filtration facilities, initial studies were begun by
Pettet et al. [92,93] using pilot-scale equipment.  Treatment at that
time consisted of primary settling, activated sludge and trickling
filters in series, and final settling.  The investigations were con-
ducted using two pilot-scale pressure filters operating at a hydrau-
lic loading of 2 gpm/sq ft, which filtered plant effluent over an 11-
month period.  Filter housings consisted of vertical, cast iron cyl-
inders 3 ft high and 2 ft in diameter capped with a flanged dome and
seated with a flanged cone.  The overall bed depth in each filter was
24 in.; however, one was equipped with sand graded 0.85 to 1.7 mm and
the other with 10 in. of 2.1 to 6.0-mm anthracite under 14 in. of
0.85 to 2.1-mm anthracite.

Backwashing was scheduled on a 24-hr interval and was applied at a
rate of 14 gpm/sq ft using filter effluent.  Spent washwater, com-
prising approximately 3% (300 gal.) of the throughput, was discharged
through a valve in each housing and returned to the treatment plant
settling tank.  To assist in the breakup of the clogged bed, each
filter housing was equipped with a handwheel-operated rake whose
teeth extended 1 ft below the media surface.  As the study progressed,
daily head loss increased from 4.5 to 30 ft because of inefficient
cleaning.  An examination of the filter  interior revealed that the
filter walls and sand were covered with  gelatinous coatings of 1/2
in. and 1/4 in., respectively.  Sodium hypochlorite solutions were
applied in varying doses and for several contact times, the most
successful combination being 0.05% sodium hypochlorite as chlorine
with a 2-hr contact  time.  After this cleanup procedure, no addition-
al chlorination was  required.

Immediately following the conclusion of  the first Luton study, Pettet
et al. [92,93] began a  second pilot-scale  study at  a nearby trickling
filter plant,  the Finham Works in Coventry.  Again, separate  sand and
anthracite filters were investigated except that they were now grav-
ity operated.  One 12-in. diameter housing contained 24 in. of sand
graded 1.0 to  2.0 mm and the other contained 3-in.  of 2.6  to  6.0-mm
anthracite supporting 24 in. of coal graded 1.0 to  2.0 mm.
                                   79

-------
Backwashing  requirements were  approximately  27=, of  the  filter  through-
put,  and each bed was  air  scoured prior  to introducing the washwater.

During  the latter four months  of the Finham  study  the  terminal head
loss  of the  sand filter approached 8 ft  on occasion where it  normally
did not exceed 2 ft.   This problem was reduced by  using approximately
three times  the normal washwater volume  during the backwashing fol-
lowing  every third run.  At no time was  chlorination required to con-
trol  gelatinous accumulations  within the filters.  Based upon the
data  available from both studies Pettet  concluded  that air scour was
worthy  of additional study as  a backwashing  aid.

Upon  completion of the Finham  study, Pettet  et al. [93,94] returned
to the  Luton Corporation Works to conduct additional experiments with
pressure and gravity filtration.  Equipment  for the pressure  filter
study were the same as previously described  for Luton.  Media type,
depth,  and gradation were changed several times during the study.
Initially one filter contained 24 in. of 0.35 to 1.7-mm sand  and the
other,  24 in. of 1 to  2-mm anthracite.   These two  filters were oper-
ated  at rates varying  from 1.2 to 4.0 gpm/sq ft over a seven-month
period.  Backwashing of each filter was  conducted  on a 24-hour basis
and was preceded by three minutes of air scour at  18.3 scfm (5.8
scfm/sq ft).  Washwater was drawn from the filter  effluent tank and
pumped  at rates of 14  gpm/sq ft and 8 to 10  gpm/sq ft  for the sand
and coal filters, respectively.  Total washwater volume was approxi-
mately  2 to  3% of throughput.  Throughout the comparison of sand and
coal  filters, no appreciable differences in  effluent quality  were ob-
served, but  effluents  from each deteriorated markedly  after six
months  of operation.   An inspection of the media in both filters re-
vealed  that  the individual particles were heavily  coated with a bio-
logical slime which was believed to be the source  of the problem.
Pettet  et al. [93,94]  considered this the result of inadequate air
scour,  so the media was chlorinated and, subsequently,  the air-scour
time  was increased to  10 to 15 min.  The effluent  quality of  both
filters was  restored to previous levels  and  the increased air scour
appeared to maintain the cleanliness of  the  media  so that no  addi-
tional  chlorination was required.

The next phase of the  study was directed at  determining the effect of
depth on effluent quality.  The old media in each  filter was  replaced
with  a  single size range (0.85 to 2.06 mm) of hand-graded sand placed
in one  filter to a depth of 24 in. and in the second to a depth of
42 in.  It was thought that the deeper bed would prevent solids
breakthrough at higher hydraulic loadings, but it was  discovered that
the shallow bed reached terminal head loss before  exhibiting  break-
through.  The filters were operated for  two  months at  hydraulic load-
ings up to 5.7 gpm/sq  ft, but no appreciable differences in effluent
quality were observed  below loadings of  4.6  gpm/sq ft  and only slight
improvement by the deep filter at rates  above this figure.  Backwash
requirements for the shallow sand bed were the same as those  for the
sand  bed of  the preceding phase:  14 gpm/sq  ft and 2 to 3% of the
                                  80

-------
filtered volume.  For proper cleaning of the deeper bed, however, the
wash rate had to be increased to 20 gpm/sq ft, and the total volume
of washwater used also increased [93,94].  Each was air scoured at an
unspecified rate and duration prior to the water backwash.

Based upon the data collected by Pettet et al. [92,93,94] in the ex-
tensive experimental studies previously discussed, six rectangular
rapid sand filters were installed at Luton in 1951 [41,42].  Each
filter was operated at a hydraulic loading of 4 gpm/sq ft, which was
maintained by a rate controller on the filter effluent line.  The
media in each filter consisted of 36 in. of sand graded so that 9070
of the grains were within the 0.85 to 1.67-mm size range.  Water
backwashing was preceded by air scour at a rate of 1 scfm/sq ft for
an unspecified duration.  Washwater was then pumped at a  rate of
14 gpm/sq ft from a filtered water storage tank, each filter normally
requiring a washwater volume equal to 2-1/2% of its throughput.
Spent washwater (approximately 15,000 gal. per filter backwash) was
collected in a second storage tank and fed back gradually to the head
of the plant.  The filters ordinarily exhibited an initial head loss
of 1 to 1.5 ft following backwashing and were allowed to  attain a
head of 8 to 9 ft before cleaning was again initiated.  Under average
conditions the filters required backwashing twice daily  [41,42,86,94].

In 1954 the -filter plant capacity was increased by 50% with the  addi-
tion of three more units, which corresponded to a simultaneous 507<,
expansion of the  activated sludge treatment units.  The  new filters
were essentially  the  same as the original units although the sand was
apparently graded somewhat finer.

Data gathered over the years since the installation of  the units has
shown  that the  rapid  sand filters generally performed well although
they were susceptible to effluent deterioration from poor backwashing
or  shock  loads  of sewage containing high  suspended  solids concentra-
tion  [86].  It  had also been observed that media was being lost  as  a
result of  slime coatings which decreased  the  effective  particle  den-
sity  and, thus  the grains were more easily  carried  away with the
 spent  washwater.   Backwashing procedures  were  changed  to those  as
presented in Table 5. Because of these difficulties  and of more
 stringent effluent standards recently imposed  upon  the  Luton plant,  a
new study was  initiated.

The objectives  of the investigation were to directly  compare  a pilot-
 scale Immedium upflow filter with one of the existing rapid  sand fil-
 ters  and to determine the  best means  of  backwashing each.  The re-
 sults of the  study demonstrated the  superiority of  the upflow unit
 for the following reasons  [86] :

 1.    Suspended solids removals were  consistently better even at flow
      rates 50% higher.
                                   81

-------
 2.   Greater utilization of the bed for solids removal resulted in
      longer runs and greater throughput.

 3.   High influent suspended solids concentrations did not cause
      severe upsets and were readily handled at slightly reduced load-
      ings (3.4 gpm/sq ft).

 4.   Backwash can be conducted readily with filter influent.

 5.   Reduced capital costs  resulting from elimination of high  filter
      walls (no static head  requirements).

 6.   Less down time for backwashing.

 Backwashing procedures are  summarized in Tables 5A and 6A,  which are
 copied from the original references.   The pulsating air scour  is re-
 portedly very effective in  breaking up the agglomerated media  of the
 upflow filter following a run.   Apparently,  the agitation is quite
 severe,  and the disturbance travels upward through the bed  in  a wave-
 like  motion.

 A visit  to the Luton works  in November,  1975,  revealed the  following
 developments.   Sand in the  rapid sand filter after backwashing had
 heavy organic slimes surrounding the  sand  grains,  but no mud balls
 were  visible.   The slimes in the upper sand  layers after backwashing
 caused about  12% dry weight loss when a  sample  was ignited  in  a stan-
 dard  loss  in  ignition solids  test.  The  backwash procedure  had been
 changed  to that shown in Table  5B incorporating a  pulsed air scour.
 Sand  loss  was  still a problem,  and  periodically sand  is  added  to
 maintain desired sand depth.  The dirty  sand did not  appear to affect
 the filtrate  quality.   Run  lengths  were  only about 7  hr  due to high
 influent suspended solids (about 35 mg/1)  coming from overloaded
 final  settling tanks.

 The upflow filters  have  also  experienced sand loss, which at times
 has completely exposed  the  hold  down  grid.  During high  flows,  this
 has resulted  in uplifting of  the bed  and breakthrough of solids at
 head losses of greater  than 6 ft.   The backwash procedure had  been
modified in 1975  as  shown in Table  6B  to attempt to improve backwash
 effectiveness.  The  change  included an increase in the high rate wash
 from 15  to  19  gpm/sq  ft.  The condition  of the  deep media was  not
 observed due to difficulty  of sampling.

 Immedium filter  studies.  The favorable  results  obtained at Luton
 sparked  interest  in  the Immedium upflow  filter  as  a wastewater filter
 and spawned three  additional  studies over  the next four  to  five  years.
The first of these  studies was conducted by Woods  et  al.  [142]   using
 a pilot-scale unit  at the West Hertfordshire Works and activated
 sludge plant effluent.  Details  of media sizing  and bed  depth were
not presented; however, the pilot Immedium filter was described  as a
                                  82

-------
Table 5A.  Manual backwashing procedure used on full-scale rapid
           sand filters at Luton [86],a


Minutes after
 stoppage of                  Description of operations
    inflow


      0           Inflow stopped.  Drain down commenced and, if
                  possible, continued to completely drain down the
                  water through the filter sand into the effluent
                  channel.

     15           Backwash outlet valve opened to remove any water
                  remaining above the level of the backwash weir.

     20           High-rate air scour and slow-rate backwash com-
                  menced together.

     30           Air scour stopped.  High-rate backwash commenced.

     40           High-rate backwash stopped.  Normal inflow
                  commenced.


filtered water used for backwashing, water and air rates were not
 specified.
Table 5B.  Backwashing procedure in full-scale rapid sand filters
           at Luton in November 1975.


 Minutes  after
inflow  stoppage                        Operation


      0-28             Drain down to bottom of sand level.

     28-32             Low-rate air and water backwash simultane-
                       ously to a water level 9 cm above the fixed
                       bed  surface.  Air at 1.23 cfm/sq ft and
                       water at 2.1 gpm/sq ft

     32-41             Intermittant air scour.  Air on at 1.23
                       cfm/sq  ft for 45 sec and off for 45 sec, and
                       cycle repeated  for 9 min

     41-49             High-rate backwash at  13.7 gpm/sq ft
                                  83

-------
Table 6A.  Automatic backwashing sequence used on pilot-scale
           Immedium upflow filter at Luton  [86].a
 Minutes after
 initiation of
backwash cycle
                         Description of operations
 0
19
26
                   Backwash cycle initiated by pressure switch.
                   Inflow stopped.  Drain down commenced.

                   Drain down completed.  High rate air scour (4.8
                   cfm/sq ft) and slow-rate backwash (2.4 gpm/sq ft)
                   commenced together.

                   Slow rate backwash stopped by electrode 9 in.
                   above sand level.  Pulsating air-scour commenced
                   (3.3 cfm/sq ft for 45 sec alternating with 0.5
                   cfm/sq ft for 45 sec).

                   Pulsating air scour stopped.  High-rate backwash
                   (15 gpm/sq ft) for 30 sec, then dropped to slow
                   rate backwash for 10 sec, then increased to high-
                   rate backwash.  (This pulsing of the backwash
                   rates ensures that any entrapped air bubbles
                   rise up out of the sand bed.)

                   High rate backwash stopped.  Normal inflow
                   c ommenc ed .
      water used for backwashing.
Table 6B.  Backwash procedure for full-scale upflow filters at
           Luton in November 1975.
 Minutes after
inflow stoppage
                                Operation
      0-27

     27-31



     31-37


     37-47
              Drain down to bottom of sand.

              Air and water wash together to a level 9 cm
              above sand surface.   Air at 2.85 cfm/sq ft and
              water at 2.5 gpm/sq ft

              Intermittent air scour at 2.14 cfm/sq ft, 45
              sec on,.45 sec off for 6 min

              High rate backwash at 19 gpm/sq ft
                                 84

-------
typical unit, which would mean that it consisted of a 1-ft layer of
gravel topped by 5 ft of 1 to 2-mm sand.

Operational advantages of the unit were:  excellent ability to handle
hydraulic shock loads, high suspended solids and biochemical oxygen
demand (BOD) efficiency at higher flow rates per unit area, and ease
of backwashing.  The latter item was accomplished by draining down
the filter to the media surface and air scouring at unspecified rate,
then applying filter influent at a rate sufficient to break the com-
pression arches in the bed and fluidize the media.  This backwashing
procedure was a departure from that recommended by Boby and Alpe [13]
because the backwash water was introduced after the air scour.  No
leaves, paper, or fat was observed lodged in the base of the filter,
which is the point of initial contact with the influent and backwash
water [142].  The backwash water removed 1.33 Ib  of entrapped solids
per square foot of filter area and required an average 1056 gal. for
each wash.

At approximately the same time, Truesdale and Birkbeck [131] were
conducting a similar Immedium filter study at another activated
sludge plant in Letchworth.  Their pilot-scale apparatus consisted of
a 2.5-ft diameter housing containing a 1-ft layer of graded support
gravel and 5 ft of 1 to 2-mm sand.  The media-restraining grid con-
sisted of a series of parallel bars at 4-in. centers located 2 in.
below the sand surface.  When the head loss reached a predetermined
value, cleaning was initiated with an air scour at unspecified rate
followed by backwash with filter influent at 14 gpm/sq ft for 15 min.
The total volume of washwater used averaged 5.7% of the throughput.

Michaelson [83] conducted a third study on upflow filtration using
trickling filter effluent from the plant at Ashton-Under-Lyne.  Few
details about the backwash were presented except that backwash was on
a daily basis and used 1.25% of the throughput.

Derby.  An extensive, pilot-scale investigation of wastewater filtra-
tion was begun in 1966 by Joslin and Greene [68] using trickling fil-
ter effluent from the treatment works at Derby.  During the next two
years, seven separate filters using three different flow configura-
tions were studied:

1.   A 24-in. deep downflow sand filter with media graded 1.2 to
     2.4 mm.

2.   A 36-in. deep downflow sand filter with media graded 1.2 to
     2.4 mm.

3.   A 24-in. deep downflow sand filter with media graded 1.2 to
      1.7 mm.
                                   85

-------
4.   A 24-in. deep triple-media downflow filter consisting of anthra-
     cite, sand, and garnet graded 1.4 to 2.4 mm, 1.2 to 1.4 mm, and
     0.71 to 0.85 mm, respectively.

5.   A 24-in. deep upflow sand filter graded 0.71 to 2.40 mm.

6.   An Immedium upflow sand filter with 26 in. of graded gravel and
     48 in. of 1 to 2-mm sand.

7.   A radial flow, moving bed sand filter containing 0.5 to 1.0-mm
     sand.

Joslin and Greene demonstrated that with respect to filtration of
Derby effluent, flow direction was insignificant in comparison to
proper media depth and gradation.  This was in direct contrast to
several other studies where the Immedium upflow filter had demon-
strated superiority to conventional downflow sand filters [83,131,
142].  Although flow configuration appeared to play no role in im-
proved suspended solids removal, increased bed depth did improve re-
moval, as evidenced by the superior efficiency of both the Immedium
upflow (4-ft depth) and deep-bed rapid sand filters (3-ft depth).
Head loss development was less rapid in the filters containing 1 to
2-mm graded sand than those, such as the triple-media filter, con-
taining finer particles.  Because of the subsequently longer filter
runs and the ability to meet desired effluent standards at high
loading rates, the overall conclusion of the study was that deep-bed
filters of coarse 1 to 2-mm sand were most suitable for a full-scale
application.

Backwashing experiences with the filters at the Derby installation
were interesting and were reported in some detail.  An inadequate air
supply precluded the use of air scour prior to applying washwater to
the media, so an attempt was made to compensate for this deficiency
by backwashing at higher rates.  Difficulties were encountered in all
the units, however, with mud ball and agglomerate formations which
resisted breakup by the water (only) backwash.  This problem was par-
tially relieved by draining the water level to within 1 in. of the
surface and then directing a jet of water onto the surface of the
filter prior to backwashing.  Mud balls which escaped breakup by the
jet tended to settle deeper in the filter.  Difficulties were also
encountered at the start of each backwash because the filter media
tended to rise as a plug upon initial application of the washwater
and caused fluctuation on the backwash rate indicator.

Minworth.  Activated sludge effluent from the Minworth treatment
plant was used as filter influent in an investigation conducted by
Tebbutt [129].  Three pilot-scale filters were assembled.  Total bed
depth in each filter was 24 in., with the first filter containing 0.5
to 1.0-mm sand, the second filter containing 1.0 to 2.5-mm anthracite,
and the third containing half 0.5 to 1.0-mm sand and half 1.0 to 2.5
mm anthracite.  Flow rates to each filter were varied from 1.7 to
                                 86

-------
10.3 gpm/sq ft, and no substantial drop in efficiency with increasing
rate was observed.

A second series of in-plant tests was also conducted using 24-in. bed
depths, the three filters having one of the following specifications:
1.0 to 2.5-mm anthracite, 1.2 to 2.4-mm sand, 2.4 to 4.7-mm sand.
Flow rates applied to the filters were the same as for the first in-
plant series.

Backwashing details provided in the text of the article by Tebbutt
are scanty, but no air scour was used, and difficulties were encoun-
tered with the media rising as a "plug" at the start of the wash, a
problem also reported by Joslin and Greene [68].  Filter beds con-
taining anthracite were most pronounced in exhibiting this problem,
while the sand beds were generally easier to maintain.  Coarse sand
(1.2 to 2.4 mm) was not expanded but did clean readily.  Tebbutt felt
that the rising plug problem could be eliminated in a full-scale
plant with the use of air scour and that this would place the anthra-
cite media in an operating advantage since the Minworth studies
showed less washwater was required for anthracite filters in spite of
the plugging.

South African Studies

Ancor.  In 1947 Vosloo [136] reported the results of an investigation
in which a pilot-scale pressure filter was used to treat the effluent
from the Ancor waste disposal plant.  Although the nature of the
treatment of the Ancor plant was not described by Vosloo, Huang [59]
described it as having secondary treatment.  A reasonable assumption,
considering the climate and other installations in the area [87], is
that the plant consisted of primary settling, trickling filters, and
final clarifiers.

The pilot filter was contained in a 2-ft diameter steel casing having
a height of 4.5 ft.  The filter media consisted of a 24-in. layer of
graded 0.5-mm sand and 5 in. of 0.8 to 1.7-mm sand, both layers of
which were supported by 7 in. of graded gravel.  Flow configuration
was downward, and effluent from the filter was collected by nozzle
mounted in a steel plate false bottom.  In a manner similar to the
arrangement on the early Luton pressure filters [92], the housing for
the Ancor filter was equipped with a handwheel-operated rake to break
up the media surface during backwashing.  Unfortunately, no other de-
tails about backwashing effectiveness or backwashing problems were
presented.

Pretoria.  Since  1954 full  scale gravity filters have been used at
the Pretoria Purification Works to remove suspended solids from a 3
mgd flow at  the trickling filter plant [87].  The five filters are
each 13.5 ft in diameter and can develop a maximum head of 9 ft.  Air-
scour capability  for backwashing has been provided on the four filter
beds equipped with a steel plate false bottom containing 1-in.
                                   87

-------
diameter nozzles at 6-in. centers  [72,88].  Compressed air is intro-
duced below the false bottom of these filters by a 1-1/2-in. diameter,
T-shaped head and is distributed through the nozzles to the media.
The remaining filter is equipped with a United States-manufactured
carborundum block false bottom and rotary surface washers.  Each fil-
ter contains sand media graded to 0.55 to 0.85 mm for 21 in. and 0.85
to 3.2 mm for 6 in., and those equipped with the steel plate (nozzle)
underdrain have been provided with a 12-in. layer of support gravel.

Designed for a maximum rate of 3 gpm/sq ft, the filters have been
operated at an average rate of 2.6 gpm/sq ft to a terminal head loss
of 6.5 ft  [72],  During an average 8 to 10-hr run, the filters con-
sistently reduced the influent suspended solids level of 22 mg/1 from
the treatment works to less than 5 mg/1 [102].  Backwash pumps were
sized to provide a maximum rate of 45.6 gpm/sq ft, which is suffi-
cient to expand the sand 50%; however, a rate of 25.7 gpm/sq ft has
been found to be adequate [ 111].  Washwater volume normally comprised
192,000 gal. or approximately 9.5% of the treated flow.  Interesting-
ly, Nicolle [87] has reported that the filter equipped with rotary
surface washers had slightly longer runs and appeared quite clean
compared to the filters equipped with air scour.  Several comments by
Nicolle [87,88] with respect to design and operation of the filters
are important:

1.   Filter influent at Pretoria is chlorinated at a dosage 4.7 mg/1
     to control biological growths and deposits and is considered
     highly beneficial.

2.   The filters have demonstrated a distinct susceptibility to hy-
     draulic shock caused by stopping, starting, or altering the
     application rate and have required backwashing soon thereafter.

3.   The use of a circular shape for wastewater filter housings should
     be avoided because of increased space requirements and diffi-
     culty in altering washwater trough placement.
                                  88

-------
          VIII.  EXPERIMENTAL COMPARISON OF BACKWASH METHODS

                       IN WASTEWATER FILTRATION


                         Pilot-Plant Equipment

The Ames Plant

Experimental work comparing different backwashing methods was con-
ducted at the Ames Water Pollution Control Plant using a pilot-scale
filter plant designed and built for this study.  The Ames treatment
plant facility employs biological treatment via three standard rate
trickling filters.  The plant was completed in 1951 and at the time of
this study (1973-75) was somewhat overloaded.  Raw sewage enters the
plant through a comminutor pit and then is lifted to an aerated grit
removal chamber.  Effluent from the grit chamber is then passed
through four rectangular primary sedimentation tanks which provide
about 2 hr detention time at a flow rate of 4 mgd.  The primary tank
effluent is then applied to the trickling filters by rotary distri-
butors and thence to three circular final settling tanks.  Although
chlorination facilities have been provided, they have never been used
because of hydraulic difficulties in discharging from the tank to the
Skunk River.  Therefore  after final clarification, the flow is col-
lected and discharged to the Skunk River.

Primary sludge and  scum  are collected and directed to an anaerobic
digester.  Final sludge  is drawn off from each of the tanks and mixed
with raw sewage  at  the head of the plant.

Serving both the University and the city, the plant treats sewage
which is primarily  domestic in nature and not subject to unusual load
or  strength variations from large industrial water users.  Variations
are observed in  sewage strength, however, during the regular  school
year and the summer vacation, when the university population  fluc-
tuates widely.  Furthermore,  the sewage collection system is  quite
susceptible  to infiltration  and inflow,  and high flows of dilute
sewage  are not unusual during wet weather.

General Arrangement

The pilot plant  was arranged  as shown in Fig.  23 for the influent,
flow-splitting method of rate control.   In  this arrangement  the in-
fluent  to each of  the three  filters  was  provided by  throttling  the
supply  pump  discharge and then  splitting the pump discharge  equally
into  thirds.   Outlet lines from the  filter  housings were connected
to  a  single  effluent line which was  placed  at  an elevation  above  the
surface of the media in  the  filters  and  defined the  low  static
water  level  in the  filter boxes.   As head  loss developed  in the fil-
ters during  the course of a  run,  the water level  above  the  media  sur-
face continued to  rise until it reached an elevation equal  to the
                                   89

-------
            FLOW SPLITTER
   .ROTAMETER-^


     SPLITTER BOX
     BYPASS LINE
BACK-
WASH
WASTE
NORTH
FILTER
                                       ROTAMETER
                            -ex	1    | SUPPLY
                                         I PUMP
BACKWASH
 WASTE
SOUTH
FILTER
                                                SURFACE
                                                WASHER
                                                SUPPLY
   r
BACKWASH
 WASTE
WEST
FILTER
EFFLUENT
 PUMP
                                    _J*    ?
                                    *	1  i     I
            FLOWMETER
                           PRESSURE
                           GAUGE
                                              ROTAMETER
          BACKWASH PUMP
                                           COMPRESSED AIR
BACKWASH
 STORAGE
  TANK
      PILOT PLANT EFFLUENT
Fig. 23.  Schematic representation of pilot-scale filter plant
         used in experimental investigation.
                            90

-------
height of the bypass line in the influent flow-splitting device.
This latter elevation defined the high static water level, or termi-
nal head loss, which was available in the pilot plant.  When filter
head loss reached the terminal stage, part of the influent was by-
passed automatically, and the filter ceased to operate at a constant
and equal rate and thereafter declined in rate until it was back-
washed.

                       Chronology of Experiments

The experimental work with the pilot plant was conducted in several
phases in which different filter media and or backwashing techniques
were used.  The phases are summarized in Table 7.  In addition to the
differences shown in the table, there was also a difference in source
of backwash water.  During Phases I and II the filtered effluent was
used as the backwash supply by passing it through the backwash stor-
age tank of Fig. 23.  In Phases III through VI, the unfiltered sec-
ondary effluent was used as a backwash source.

                           Equipment Details

The west filter was completed during the middle of Phase I.  Since
a complete set of data were not collected for that filter on Phase I,
the data for  that filter in Phase I are not presented herein.

Flow Splitting

Two different flow splitters were used.  During the first half of
Phase I, a standard 1-in. pipe tee was used.  Globe valves on either
branch were used to adjust the split.  This arrangement was easy to
construct and functioned reasonably well.  During the last part of
Phase I, it was desired to split the flow three ways  so that the West
filter could  be used.  This was accomplished by use of an influent
splitter box  which was used for all subsequent Phases.  The splitter
box consisted of two fabricated steel boxes, or halves, that fit on
top of each other.  The influent flow was split equally to the three
filters by identical orifices  located in the bottom of the upper box.
A float valve on the influent  pipe was used to maintain a constant
water  level on the orifices in the upper box.  Discharge  from each of
the three orifices dropped through an air gap into one of the inlet
lines  to  the  filters which were connected to the bottom of the lower
box.  When one or more  of  the  filters reached terminal head loss,
water would back up  in  the lower box  and overflow to  waste.  The  fil-
ter inlets and overflow line were arranged  so that the other filters
continued  to  function without  interruption  or change  in flow rate.
This  arrangement  allowed completely  independent  operation between  the
three  filters.   Several sets of  identical orifices were available
which  could be  inserted in the flow-splitter box to  achieve various
desired  filtration  rates.  The actual  filtration rate to  each  filter
was measured  with  a  rotameter  on  the  overflow  line.   The  flow  to any
                                   91

-------
Table 7.  Summary of experimental phases for wastewater filtration backwashing study.
Phase
I
II
III, IV
and V
VI
Period
5/17/73-
8/21/73
8/28/73-
10/20/73
5/16/74-
ll/ 2 Ilk
6/4/75-
8/4/75
Filter
Influent
Alum
treated
sec. effl.
Secondary
effl.
Secondary
effl.
Secondary
effl.
Media
Na sa wa
Identical dual
media
Same
Dual Coarse Mixed
media sand (tri)
media media
Identical coarse
sand

N
Water
only
n
Air
scour
type 2b
Backwash
S
Air
scour
type lb
n
Air and
water
together
type 3b
Identical air and
type 3b

W
Fixed
surface
wash
Rotary
surface
wash
Surface
and
subsurface
wash
water
&N, S, W refer to the north, south,  and west filters, respectively.
 Type 1 air scour means air first  followed by water with full-bed fluidization and expansion.
 Type 2 air scour means air and water  together briefly during the rising water level followed by
 water only with full-bed fluidization and expansion.
 Type 3 air scour means air and water  together during most of the backwash while overflow was
 occurring.  Used only for coarse  sand media which was not fluidized or expanded during the backwash.

-------
filter was purposely bypassed to the rotameter to measure the rate
whenever desired.

Alum Treatment Equipment

During Phase I, the settled secondary effluent was pumped to an upflow
solids contact unit where it was treated with alum for phosphate
precipitation and settled before filtration.  Figure 24 shows the
upflow solids contact unit.  This unit was mounted in a truck and
formerly served as part of a mobile water purification plant for the
United States Army.  The conical shaped settling and reaction tank is
referred to as an erdlator.  It had a volume of approximately 530
gal.

Raw water was pumped through a pair of nozzles into two troughs.
This aerated the water just before it overflowed into a vertical,
cylindrical, center column in the erdlator.  Alum solution was added
in one of these troughs.  The center well contained an agitator that
mixed the alum solution with the raw water and provided some floccula-
tion.  Rapid mixing was not provided.  After flowing downward through
the center column, the water moved upward to a collection trough.

The flocculant precipitate comprised of A1P04 and A1(OH>3 separated
in a floe blanket where the upward water velocity became equal to the
hindered settling velocity of the floe.  The blanket of floe provided
an environment that aided the reactions with the alum and encouraged
flocculation or the agglomeration of smaller floe particles into
larger ones.  Either sedimentation or enmeshment in the floe blanket
removed much of the material suspended in the erdlator influent.  The
surface of the floe blanket was easily observed through the clear
water above.  The blanket itself appeared to be fairly thick, but its
lower surface could not be observed.

The clear water above the floe blanket was collected in a circular
trough.  Orifices on both sides near the top of the trough provided
for even collection of water.  From the collection trough the water
flowed to a clear well.

As the floe in the erdlator formed and as sewage particles were en-
meshed in it, a considerable volume of sludge was created.  A slot in
the side of the erdlator drew the sludge from the top of the blanket
into a sludge concentrator tank.  The clear supernatent from the con-
centrator was drawn off into the clear well and the concentrated
sludge was drawn to waste by a positive displacement pump.

A dual head, diaphram type, positive displacement pump was used to
feed the alum solution.  The pump drew from two 150-liter containers
where the alum was prepared.

Water was drawn from the clear well by a self-priming centrifugal
pump.  A rotameter and a globe valve in the pump discharge line
                                  93

-------
                                             COLLECTION
                                              TROUGH
                                           TO POP FLOC
                                            BLANKET
                                           15 in.
 UPFLOW
 SETTLING
 ZONE
 39 In.
                                        CENTER
                                        COLUMN
                                      MIXING ZONE
CENTER
COLUMN
COLLECTION
TROUGH
       Fig. 24.  Pilot plant solids contact unit.
                           94

-------
enabled quick flow measurement and adjustment.  The pump then dis-
charged to the flow splitter.

Filters

Housing details.  Each of the three filter housings used in this ex-
perimental investigation was fabricated from 3/16-in. steel plate and
was equipped with a 1/2-in. thick plexiglass front for viewing the
interior during the backwashing cycle.  Vent pipes were installed on
each housing to permit the filters to operate by gravity for this
study, although each unit can be completely sealed for pressure fil-
ter operation.  The interior horizontal cross sectional area of each
filter was 2.25 sq ft, and other pertinent interior dimensions are
shown in Fig. 25.  Access to the interior of the housing was provided
by a top mounted, 6-in. diameter, hand hole.  At the beginning of
Phase VI, all three filter housings were extended in height to 10 ft
so that deeper media could be used.

Located near the top of each housing was a 1-1/2 by 4-in. inverted
bell mouth which distributed the influent during the course of a fil-
ter run and collected the washwater during the backwash cycle.

Each filter housing was equipped with an underdrain system consisting
of a steel plate false bottom and five General Filter Company media-
retaining strainers (or nozzles).  During a normal run the nozzles
collected the filtrate and directed it to the plenum below the false
bottom where the effluent lines carried it to the backwash storage
tank.  When cleaning was required, water and air (if used) entered
the plenum and were distributed uniformly to the media through the
nozzles.  The media-retaining nozzles had openings of 0.4 mm during
Phases 1 and II.  There was evidence of partial clogging of these fine
openings by the end of Phase II.  Therefore, nozzles with 1-mm slots
were installed at the beginning of Phase III and graded gravel was
needed for the north and west filters to prevent loss of filter media.
A nozzle clogging incident occurred during Phase III which required
termination of that phase.  During rebuilding of the two filters, new
nozzles with 4.5-mm slots were installed.  The incident will be de-
scribed later in the results section.  Nozzles with 4.5-mm slots
were used in all three filters in Phase VI.

Head loss development in the filters was monitored by seven piezom-
eters attached to taps located vertically along one side of the fil-
ter housing.  The taps were located at 4-in. centers starting just
above the underdrain plate.  Each tap consisted of a 3-in. length of
1/4-in. copper tubing soldered to 1/4-in. brass fittings which could
be threaded directly into  the side of the housing.  The  interior end
of the tap extended slightly into the media  and was covered by a fine
stainless steel mesh.  The exterior end of each tap was  connected to
a length of 1/4-in. clear  plastic tubing which was then  mounted to  a
piezometer board.  Thus, using the false bottom as a datum, the
pressure at several points within the filter media was determined by
                                   95

-------


—
^s
\
-«-

—
•J
•^ —
-^ —
O—-0

1 ff 6 \nT
-*-PIEZOMETER BOARD
     UNDERDRAIN PLATE

             |10 1/2 in
            o     o
                o
            o
          STRAINERS
    PIEZOMETER TUBES
   •PIEZOMETER TAPS
                                VENT PIPE
4:
»o
                                  in.

     •AIR INLET
     •FILTRATE OUTLET AND BACKWASH INLET
     •DRAIN AND SAMPLING POINT
                                    — 1 ft 6 in.-j
  Fig. 25.  Details of filter boxes,
                  96

-------
observing the water level in each of the clear plastic tubes.   Dif-
ferent piezometer tap positions were used in the various phases de-
pending upon the depth of media in service.

Surface washer.  The west filter was chosen for installation of a
rotating surface washer consisting of an inverted lawn sprinkler
which threaded into a coupling on the back plate of the filter
housing.  Water was supplied to the surface washer from a yard hy-
drant supplied by the treatment plant non-potable well water system.
Hydrant line pressure was usually 60 to 70 psig and surface washer
operating pressure generally was from 40 to 45 psig.

After about 30 filter runs in Phase II using the improvised rotary
surface washer, a two-armed model incorporating two nozzles was in-
stalled.  The new rotating surface washer more closely resembled
systems in actual use and produced strong, highly directional jets.
The two arms of the surface washer consisted of 1/2-in. diameter
brass nipples, were each 6-in. long, and were positioned 180° apart
in a horizontal plane.  One 1/8-in. diameter nozzle (Leopold Corpora-
tion) was placed on a 45° elbow at the end of each arm  and positioned
toward the media surface at an angle of 15° to the horizontal.

This system was used essentially as described for the remainder of
Phase II except for a slight shortening of the rotating arms to
better direct  the end nozzles to the media.  At the beginning of
Phase III, a  subsurface washer was  added to the west  filter, and  a
booster pump was added to the surface wash supply line, which  in-
creased the operating pressure to about 80 psig.  Both  nozzles were
positioned down at  a 15° angle to the horizontal for  the  surface
washer while  the subsurface washer had  one nozzle pointed up and  one
down at the same angle.  Two 1/2-in. pipes in  the filter  supported
the washers.   At the start of Series IV, both  surface washers were
modified.  Both washer arms were changed to 1/4-in. nipples  approxi-
mately 5 in.  in length.  A 1/4-in., 90° elbow was  attached  to  the
end of each arm, and a Leopold nozzle was  fastened  on the end  of  each
elbow.  Both  nozzles of  the  surface washer were pointed down at  a
15° angle.  The subsurface washer had rubber  capped nozzles, one  up
and one down,  both  at  a  15°  angle.  The vertical position of the
washers was changed to coincide more closely with  the coal-sand  in-
terface  for  the subsurface washer,  and  the surface  of the coal for
the  surface washer.

At  the  beginning of Phase VI,  the  surface  and subsurface washer were
removed  from  the west  filter.

Filter media.  The  filter media was obtained  from several sources and
 in  several sizes  for the various  phases as summarized in Table 8.  In
Phase  I and  II,  it  was desired that the dual  media be as identical
 as  possible  in each of the  filters. To achieve this  objective the
 following precautions  were  used at the beginning of each phase when
placing new media  in the filters.   The total  volume of a particular
                                   97

-------
 media  layer  required  for  the desired  finished  depth  plus  a  3-in.
 skimming depth was determined,  and  the  required number  of bags were
 poured onto  a concrete  slab.  After the media  had been  thoroughly
 mixed,  it was split and placed  in each  housing.  Next the media was
 backwashed at a  rate  of about 15 to 20  gpm/sq  ft for 15 rain to allow
 the  fine particles to work  their way  to the  surface.  Finally, with
 the  backwash rate reduced to the point  that  the coal was just flu-
 idized, the  top  3-in. layer was removed from the housing with a si-
 phon.   The sand  layer in  Phase  I was  not  skimmed in  this fashion.

 After  the layer  had been  placed and skimmed, a core  sample  of the
 layer  was taken  from  each filter and  subjected to a  sieve analysis,
 the  results  of which  are  presented  in Table  8.

 A  similar procedure was used in Phase III and  IV to  ensure  that the
 coal and sand of the  dual-  and  triple-media  filters  were identical.
 The  only difference was that the skimming depth of each layer was
 1/2  in., in  accordance  with the suppliers recommendation.   The coarse
 sand of the  south filter  was not skimmed in  this fashion since it was
 too  coarse to fluidize  with backwash  rates that were available, and
 also because fluidization was not to  be practiced in the routine
 backwash operation of that  filter.  The splitting precautions without
 skimming were also used in  Phase VI.

 Gravel  support.  No gravel  support  layers were used  in  the  filters
 during Phases I  and II, when media  retaining underdrain strainers
 were used.   However,  the  clogging problems encountered  in the strain-
 ers  led to the use of larger openings in the strainers  in Phases III
 through VI and necessitated the use of  gravel  support in some cases.
 It was desired to test  the  stability  of a double reverse graded
 gravel during air scour to  determine  if gravel movement could be
 avoided, and to  compare that behavior with a conventional graded
 support gravel.  Therefore, at  the  beginning of Phase III,  a double
 reverse graded gravel support system  was installed in the dual-media
 filter and a regular  graded gravel  system in the mixed-media filter.
 Originally,  the  double  reverse  graded system was accidentally in-
 stalled upside down,  and  the gravel started  to mound immediately.
 The  gravel was then installed properly  in the  following manner:

                      Size  (in.)      Thickness (in.)

                 Top    1/2 X 3/4           3

                       1/4 x 1/2           1.5

                       1/4 x 1/8           1.5

                       1/8 x #10           2

                       1/4 x 1/8           3

Later, when  the  filter was  rebuilt  at the beginning  of  Phase IV, the
 underdrain gravel was modified  to accommodate  strainers with larger
                                 98

-------
                Table 8.   Filter media  details  for wastewater  filtration pilot studies.
VO
Layer Details Size Information3
Media
Phase Filter type Type
I N and S Dual Coal
Sand
II N, S, Dual Coal
and W
Sand
III N Dual Coal
Sand
W Triple Coal
Sand
Garnet
S Coarse Sand
sand
IV &V NandW Same as Phase III
VI N Coarse Sand
sand
W
S "
Effective Uniformity 90% Specific
Thickness size, coeff. finer, gravity
11. 8a 0.94 1.31 1.4 1.67
12 0.38 1.50 0.73 2.65
12a 0.92 1.52 1.5 1.67
12a 0.38 1.50 0.73 2.65
15a 1.03 1.57 2.03 1.7
9" 0.49 1.41 0.82 2.65
(Same as N filter)
(Same as N filter)
3a 0.27 1.55 0.54 4.2
46 2.0 1.52 3.6
except new media installed and coal depth changed to 17 in.
24)
47) 2.5 1.28 3.7
60)
Supplier
Carbonite Filter Corp.
Del ana, Fa.
Northern Gravel Ca.
Muscatine, la.
Carbonite Filter Corp.
Northern Gravel Co.
Neptune Microfloc Corp.
Corvalis, Or.
Neptune Microfloc Corp.


Neptune Microfloc Corp.
CX Products Corp.
Brady, Tx.

CX Products Corp.
••
               ^epth and size after placement, stratification by backwashing and skimming fines from the surface of the layer with a siphon.
                Coarse sand was not skimmed.

-------
 (4.5-mm)  slots.   The support system used for the rest of the study
 was:
                       Size (in.)      Thickness (in.)
                 Top   3/4 X 1/2            3
                       1/4 x 1/2            1.5
                       1/4 x #10            5
                       1/2 x 1/4            1.5
                       3/4 x 1/2            1.5
 The mixed-media  filter employed a normally  graded support system.
 The original  system and  the  rebuilt one  are the  following:
                      Original                     Rebuilt
      Top       3  in.  of Coarse Garnet     3 in.  of Coarse Garnet
                                               (-12+16 mesh)
               2  in.  of 1/8 X #10          3 in.  of 1/4 x #10
               2  in.  of 1/4 x 1/8          2 in.  of 1/2 x 1/4
               2  in.  of 1/4 x 1/2          2 in.  of 3/4 x 1/2
      Bottom    3  in.  of 3/4 x 1/2
 As  will be reported  later, the double reverse  graded  gravel  was not
 stable in Phases III through V, so a similar but revised support was
 used in Phase  VI in  the  north filter.  The  gradations  in that  support
 were:
                          Size (in.)        Thickness  (in.)
               Top       1/2 x 3/4                  4
                         1/4 x 1/2                  2
                         1/4 X #10                 2
                         1/4 x 1/2                  2
               Bottom    1/2 x 3/4                  2
 Backwash Supply
 A centrifugal  pump mounted on the  backwash  tank  pumped washwater to
 the  filters in Phases  I  and II.  The flow rate of  the backwash water
was measured by a type of Venturi meter.  This meter was  fabricated
 by  replacing a portion of the 1-1/2-in. diameter backwash line with a
 section of 1-in.  diameter pipe using two reducing  fittings.  Pressure
 taps on the 1-1/2-in.  and 1-in. pipes were  connected to  opposite legs
 of a mercury filled manometer.  The meter was calibrated volumetri-
 cally by measuring the rate at which water  rose  in the  filters for
                                 100

-------
different manometer readings.  In Phases III through VI, the filter
influent supply pump was used also for backwashing with unfiltered
secondary effluent.  An additional pump discharge connection was made
from the pump to the same Venturi meter described above.

Air Supply

A Speedair belt driven, dual piston compressor supplied compressed
air for the air scour sequence.  A pressure switch automatically
started the compressor when the pressure in the storage tank fell
below 40 psig.  A Fisher Model 95L pressure reducer, reduced the
storage pressure to an operating line pressure of about 8 psig.  The
amount of air flow was regulated by adjusting a globe valve and
reading the percent of full flow through a Brooks rotameter.  The
manufacturer's calibration of the rotameter at standard conditions
was accepted, but was corrected to the operating temperature and
pressure of the air supply at the rotameter.  Near the end of the
research, an approximate calibration was achieved by noting the time
for a given pressure drop on the air storage tank while the air was
being used at a constant rate and the compressor was shut off.  Ap-
propriate calculations assuming the ideal gas law was applicable
verified the manufacturer's calibration.

Samplers

In Phase I, composite samplers were used to collect samples of efflu-
ent from both filters and the erdlator (Surveyor Sampler, N-Con
Systems Co., Inc., New Rochelle, N. Y.).  During Phase II, the same
three samplers were used on the three filter effluents.  A fourth
sampler was obtained late in Phase II which was then always used on
the filter influent (the Ames plant secondary effluent).  Up until
that time, the secondary effluent samples were either grab samples or
composite samples collected manually by the Ames plant operators.

Each sampler consisted of an electric motor driven positive displace-
ment pump actuated by a timing mechanism.  The sampler contained two
arms on a timing face, one of which set the pump running time and one
of which could be set to activate the pump from 3 to 20 times per
hour.  To ensure that a fresh  sample was obtained each time the pump
was running, the waste line on the discharge side of the pump was
1/2-in. diameter while the sample line was 1/4-in. diameter.  This
meant that there was a delay before the filtrate was pumped to the
sample container, thus allowing the system to first flush itself
through the waste line.  Vents were provided in both the waste and
sample lines to prevent siphoning into or out of the polyethylene
collection containers, which were kept refrigerated in  an ordinary
domestic refrigerator  to preserve the sample during the compositing
period.  The compositing period for each phase was  somewhat different
and will be described later  in the presentation of  the  results.
                                  101

-------
 To  evaluate  the  effectiveness  of  the various backwashing  techniques,
 core  samples of  the media were taken periodically  following  the nor-
 mal backwash cycle for  a  particular filter.  The core  sampler was
 constructed  of a 4-ft length of 1-1/2-in. diameter copper tubing
 fitted with  a shear gate  at one end.  With the  gate fully open, the
 sampler was  lowered to  the desired depth in the media,  at which time
 the gate was closed and the sample removed.  Because it is well known
 that  the majority of solids removal takes place in the upper layers of
 a deep bed granular filter, the penetration of  the core sampler was
 restricted to approximately the top 12  in. of the  filter  media.
 Sometimes, the backwash water  was turned on slightly, not enough to
 fluidize or  expand the  bed, but to ease the entrance of the  sampler.
 After removal of the core, the backwash water was  turned  on  at the
 normal rate  for  about 1.5 min  to even out the surface of  the media.

                          Analysis of Samples

 The analysis of  samples was the same during all phases of the study
 except for the media abrasion  test.  The procedures are summarized
 in  the following paragraphs.

 Filter Media

 Sieve analysis.  A set  of United States standard sieves was used to
 determine the media size.  Samples were dried and  then split with a
 riffle type  sample splitter until approximately 300 g of  media were
 obtained.  The media samples were hand shaken through the sieve.s
 until 1 min  of additional hand shaking changed  the weight of media
 retained on  any  sieve by  less  than 1% of the weight on  that  sieve.
 The weights  of sieves empty and with media were measured  and recorded
 to  the nearest 0.05 g.

Media density and specific gravity.  The density of the media was
determined by a  water displacement technique with  a 100-ml pycnometer.
The procedure is presented in  more detail with  illustrative calcula-
tions in a later chapter.

Media abrasion.  The media abrasion procedures  changed several times
during the research.  However,  in all cases, the purpose  of the test
was the same:  to determine the relative amount of dirt or suspended
 solids remaining on the media  after the normal  backwash procedure
 for the filter.

The procedure in Phase  I  was as follows.  Samples were reduced in
 size  by dividing them in  half  with a riffle type sample splitter.
Then  each half was divided twice again so that  two samples, each one-
eighth of the original  sample, were obtained.   The abrasion test was
 then performed in duplicate.   When samples were taken and split, a
portion of the dirt attached to the media was undoubtedly loosened.
Therefore, as the samples of sand and coal were split, the liquid
                                  102

-------
portions were also split and the dirt present included in the calcu-
lation of dirt removed from the media.

After the sample was split, it was placed in an 800-ml beaker and
stirred at 200 rpm for 10 min with a 2-in. diameter, three-bladed
propeller (Cole Farmer Constant Speed and Torque Control Unit) im-
mersed in the media.  After stirring, the dirt was rinsed into a
graduated cylinder.  Care was taken so that no sand or coal was
transferred with the dirt.  The-media was swirled and rinsed repeat-
edly with distilled water and the dirty rinse water decanted into
the graduate cylinder, until the rinse water was clear.  The total
amount of water used to wash the media was recorded.  Suspended
solids concentration of the wash water was determined on triplicate
samples according to Standard Methods [117] by filtering an aliquot
through Whatman GF/C glass fiber filter paper.  The media samples
were then dried overnight at 103 °C and weighed.   The dirt  removed
was calculated by the following formula and results expressed as mg
of suspended solids removed per gram of filter media.

          (SS of washwater, mg/1)(Volume of washwater. 1)
                    (Weight of  filter media, g)

     =    mg SS removed
             g media

A different abrasion test procedure was used in Phases II and III
through V.  The main differences were the  deletion  of  the  splitting of
the sample on the riffle type splitter (since that was not considered
very suitable for a sample containing water) and  the extension of the
period of abrasion to 30 min  (because the media was dirtier in the
direct filtration of secondary  effluent).  A motor driven support
stand was added to rotate the beaker in the direction opposite that
of the propeller.  The other details of the procedure are as follows;

The entire core sample and any  water collected with it was put in a
2000-ml beaker and placed on  the revolving stand  (Driven by a Gerald H.
Keller Co. variable speed direct current motor).  The sample was
stirred for 30 min with a 2-in. diameter,  three-blade propeller ro-
tating at 200 rpm.  The propeller  and the  stand,  revolving in opposite
directions, were positioned so  the propeller would  traverse through
the media in a circle, abrading all  the sample.   After mixing, the
supernatant was poured into a 6000-ml flask.  The media was rinsed
repeatedly with distilled water and  the dirty rinse water decanted
into the 6000-ml  flask until  the rinse was clear.  The total amount
of washwater in the 6000-ml flask  was then recorded.  The  remainder
of the procedure was unchanged  from  that  in Phase I.

The final modification of  the abrasion test procedure was  adopted at
the beginning of Phase IV due to concern  that the 30-min abrasion was
causing excessive  abrasion of the  coal itself,  and  not merely  of  the
                                  103

-------
 solids  adhering  to  the coal.  The  abrasion period was reduced to 5
 min.  The details were as  follows:

 As before,  the core sample was placed  in  a 2000-ml beaker, but this
 time distilled water was added to  bring the level up to an easily
 read mark.  The  sample was then mixed  in  the  same way as previously
 mentioned but only  for 5 min.  The coal or sand was allowed to settle
 for 5 sec before a  100-ml  sample of  the supernatant was pipetted off
 and transferred  to  a 1000-ml volumetric flask.  The flask was then
 filled  to the mark  (10:1 dilution),  and the suspended solids test
 previously mentioned was run on this suspension.  If the media was
 very dirty, a higher dilution factor was  used to speed up the fil-
 tering  step in the  suspended solids  test.  The media was dried and
 weighed as before.   The abrasion test  result  was then figured as
 follows:

     Abrasion Test  _ (Avg SS)(Dilution factor)(Supernatant Vol)
     Result (mg/g)  ~              (Weight media)

Even though the  abrasion test procedure for Phases IV and V was
 shorter and less complex, it still provided the data needed to evalu-
 ate the relative effectiveness of  the  backwashing procedures.
Naturally, the absolute values obtained were  lower due to the shorter
 abrasion period.

Abrasion tests were  not conducted during  Phase VI.

Water Quality Analyses

During  the various  experimental phases, extensive testing of filter
 influent and effluent samples were performed  to see if the backwash-
ing methods being used had any affect  on  filter performance.

Observation filter runs in which detailed observations were recorded
and composite samples were collected were made at regular intervals.
During  these "observation runs," grab  samples of filter influent and
effluent were collected for immediate  turbidity measurement in the
field.  Composite samples were collected  over the observation run and
transported to the  laboratory for  further analysis.  Most of the
analyses were conducted there by-personnel of the Analytical Services
Laboratory of the Engineering Research Institute.  Suspended solids
analyses were done by project personnel in Phases I, II, and VI, and
by the  laboratory personnel during Phases III through V.

A typical composite  sample was subjected  to several analyses:  five-
day biochemical  oxygen demand (BOD), soluble BOD, suspended solids
 (SS), total organic  carbon (TOC),   soluble organic carbon (SOC), and
ammonia (NH3).  Periodic composite samples were also analyzed for
the following:  nitrite (N02), nitrate (N03), total Kjeldahl nitrogen,
total phosphate/ and orthophosphate.   Soluble BOD and SOC were not
                                  104

-------
analyzed in Phases I and II.  The analytical procedures were as
follows.

Suspended solids.  Suspended solids tests were conducted on filter
influent samples, filter effluent samples, and abrasion test wash-
water aliquots in accordance with Standard Methods [117].  Disposable
aluminum weighing tins and Whatman GF/C glass fiber filtering discs
were placed in a 103 °C oven for 15  min to drive  off moisture  and
then cooled and weighed to 0.1-mg precision.  The filter discs were
then placed on a Millipore filter holder which was connected to a
Millipore vacuum pump.  Under a steady vacuum of 15 in. of mercury,
the following sample amounts were filtered through the discs;

1.   Filter influent, 100 ml or 250 ml depending upon the level of SS

2.   Filter effluent, 500 ml

3.   Abrasion test washwater, 25 ml

After filtration had been completed, the discs and weighing tins were
returned to the oven and dried overnight at 103 °C.  Finally,  the
discs and tins were cooled and reweighed, and the subsequent gain  in
weight was divided by the volume filtered to yield the suspended
solids concentration in mg/1.

Turbidity.  Turbidity measurements were made using a Hach Model 2100
turbidimeter during Phase I and a Hach Model 2100 A during the re-
mainder of the studies.  Both turbidimeters have four ranges gradu-
ated in Formazin turbidity units (FTU).  Formazin standards provided
by the manufacturer were used to properly calibrate the instrument in
each range.  For the purpose of this study, only the 0 to 1, 0 to  10,
and 0 to 100 FTU ranges were required.  Glass cells were rinsed with
distilled water before  each use, and care was taken to dry the outside
of each cell before  inserting it into  the instrument.

TOG and SOC.  Total  organic carbon  (TOG)  and soluble organic carbon
(SOC) determinations were conducted on a  Beckman Model 1R315 carbon
analyzer using the tentative procedure outlined  in Standard Methods
[117].  Fifty-milliliter portions from the  filter influent and efflu-
ent samples were transferred to small  polyethylene bottles, treated
with several drops of  concentrated  112804,  and then refrigerated until
analysis.  The depressed pH, as well as the refrigeration, retarded
bacterial  decomposition of  the  organic carbon present.

At the  time of analysis, the containers were purged with nitrogen to
eliminate  carbonate  and bicarbonate interferences  (in  conjunction
with the acidification step).   Samples of 25 p.1  size were analyzed in
triplicate by comparison with acetic acid standards.   Soluble  organic
carbon  samples were  filtered through Whatman GF/C  glass  fiber paper
prior to analysis.   The carbon  analyzer was down for  repairs  for ex-
tended  periods during the course of these measurements and  the
                                  105

-------
acidified samples were held under refrigeration for long periods
while repairs were made.  This is a questionable procedure and thus
the TOG and SOC results reported herein are not considered to be of
uniform reliability.

BOD.  Determinations of BOD and soluble BOD were conducted using a
dilution technique as outlined in Standard Methods [117].  A con-
tainer of primary effluent from the Ames Water Pollution Control
Plant was supplied with each new set of samples for use in seeding
the dilutions.  Because of a misunderstanding between the laboratory
and the research personnel, nitrification was not suppressed in the
BOD analyses in Phases I and II.  Nitrification was suppressed there-
after using "N-serve" from Hach Chemical Company.  Soluble BOD sam-
ples were filtered through Whatman GF/C glass filter paper prior to
analysis.

Other analyses.  Procedures for the remaining three analyses were
automated using a Technicon Autoanalyzer.  Total Kjeldahl nitrogen,
which included both organic and ammonia nitrogen, was conducted as
described in Standard Methods [117].  The test differs from the clas-
sic Kjeldahl determination of organic nitrogen only in that ammonia
is not removed from the sample before digestion and distillation.
For use on the autoanalyzer a colorimetric technique using the tenta-
tive phenate method was employed for the final nitrogen measuring
step of the procedure.

Both phosphate determinations were also conducted in accordance with
procedures described in Standard Methods [117],  Orthophosphates were
measured using the tentative ascorbic acid procedure, which involved
a colorimetric determination preceded by filtration of the sample.
The digestion step of the total phosphate analysis was modified by
using both perchloric and sulfuric acids in the presence of a sele-
nious acid catalyst to convert the various phosphate forms to ortho-
phosphate.  A colorimetric determination using the vanadomolybdo-
phosphoric acid method was employed for final phosphate measurement
in the autoanalyzer.

                    Operation and Results - Phase I

                 Dual-Media Filtration of Alum-Treated
                           Secondary Effluent

Phase I was a comparison of two backwashing methods during the fil-
tration of alum-treated secondary effluent.  The alum treatment was
to reduce the phosphorous level of the secondary effluent to about
1 mg/1 phosphorous.  Both filters were equipped with dual media.
One was washed by water fluidization only.  The other was washed with
air scour followed by water fluidization, hereinafter referred to as
air-scour auxiliary.
                                 106

-------
Operation - Phase I

A filter run started with a backwashed filter and ended when the fil-
ter had operated for a cycle and had then been backwashed and was
ready for service again.  Filter runs were numbered consecutively
starting with number 1.  Since both filters were backwashed at the
same time, and placed in service simultaneously, a particular run
number refers to both filters.

The first filter run of Phase I started on May 17, 1973, and the last
one using alum-treated water ended on August 21.  The filters were
operated continuously during this period except during equipment
failures and when major maintenance work was needed.  The filters did
not operate for a total of about 1-1/2 weeks.

Solids contact unit.  Commercial grade, granular alum (approximately
Al2(S04>3 • 14 H20, approximate molecular weight 600) was mixed at the
rate of 37.8 g/1 of water.  This solution was fed to the erdlator to
achieve an alum dosage of 200 mg/1 when the erdlator was operating
at 10 gpm.  For runs 1 through 45, the first 7-1/2 weeks of operations
the erdlator operated at approximately 10 gpm, but for runs 46 through
57, the next 3-1/2 weeks, it was operated at 12 gpm and therefore the
alum dosage was only 167 mg/1.  For the last three weeks, after run
57, the erdlator was operated at approximately 15 gpm, and the alum
feed was increased so that the dosage was 200 mg/1.  The detention
time in the upflow portion of the erdlator was 53 min at 10 gpm and
35 min at 15 gpm.  The detention time in the center mixing column was
11 min at 10 gpm and 7 min at 15 gpm.  The surface loading was 0.37
gpm/sq ft at 10 gpm and 0.56 gpm/sq ft at 15 gpm based on the area
outside of the center column.  Sludge was wasted at a rate of about
1 gpm.

The use of alum was discontinued after run 73, 13-1/2 weeks of opera-
tion.  For another week, or eight filter runs, the trickling filter
effluent flowed through the erdlator without any chemical treatment.

Filters.  The filtration rates averaged 2 gpm/sq ft for runs 1 through
45 and for runs 59 and on.  The filters were operated at 1.6 gpm/sq ft
for runs 46 through 57.

After about 12 weeks, run number 65 for the south filter and run
number 66 for the north filter, a layer of fine sand was skimmed from
the top of the coal.  This layer was removed to see what effect it
was having on the development of head loss in the filters.

Backwash.  A combination of time and head loss was used  to determine
when the filters were backwashed.  If one of the filters approached
the maximum available head loss of 7 ft, both filters were back-
washed.  Frequently the filters were backwashed after 24 hr  of  ser-
vice even though neither filter was near the maximum head loss.
                                  107

-------
For  runs  1  through  26,  air  scour was not used on either the north or
the  south filter.   The  backwash for run 27 consisted first of the
normal water  fluidization backwash.  Samples of filter media were
taken for the abrasion  tests.  Then the south filter was air scoured.
Following the air scour, a  second normal water backwash was conducted
and  waste washwater samples were collected at various intervals of
time during the backwash.  Another sample of media from the south
filter was  then taken.  From run 27 to the end of  Phase I  (run
78), the  north filter received water backwash only, while the south
filter received air scour followed by water backwash.

After run 78,  the very  end of  Phase I, a series of special back-
washes was  used in  an effort to restore the filter media to a clean
condition.  After the normal backwash, the following procedures were
used on both  filters in the order given.

1.   A normal air scour followed by a normal water backwash.

2.   Simultaneous air and water backwash (as the water rose from the
     media  to near  the  overflow) followed by a normal water backwash.

3.   Step 1 repeated.

4.   Step 2 repeated.

5.   Step 2 repeated.

During each step samples of waste backwash water were composited for
later analysis.

The  normal  water backwash procedure at any time was the same for both
filters.  For runs  1 through 28 a constant rate of 20 gpm/sq ft for
5 min was used.  For runs 29 through 63, both filters were backwashed
for 5 min at  rates  that expanded the beds 38 to 40%.  The backwash
rates to  achieve these  expansions changed from one backwash to another
and varied  between  the  two filters.  Rates from 18 to 23.8 gpm/sq ft
were required.  After run 63, a constant backwash rate of 18.6 gpm/
sq ft was adopted,  again for a 5 min duration.  Filter bed expansion
then varied from 25  to  38%.  When air scour was used, it consisted of
9 scfm [scfm  (standard  cubic foot per minute at 14.7 psi and 70 °F)]
or 4 scfm/sq  ft of  air  for 5 min.
                    i
The backwash procedure  included the following details.  The flow to
both filters was stopped.  Then effluent valves were closed, the
valves to the  composite samplers were closed, and the backwash waste
drain valve was opened.  If air scour was to be used, the effluent
valve on  that  filter was left open until the water level had drained
at least  1  ft  below the backwash water outlet.  This was to prevent
loss of media  due to the violent agitation caused by the air scour.
After the air  scour was terminated, the water wash was started at a
very low  rate until all of the trapped air had escaped.  This usually
                                  108

-------
took 1 to 2 min.  The filter was then backwashed with water.  The
backwash water inlet valve was gradually opened over a 20-sec period.
The water backwash continued for 5 min, and then the inlet was closed
over a 10-sec period.  It was closed in as nearly as possible the
same manner each time.

After both filters were backwashed, flow was started through them
simultaneously.  Piezometer readings and turbidity measurements were
taken about 15 min later.  This allowed time for head loss and flow
to reach an equilibrium with each other.  Also any backwash water
left in the filter was flushed out by then.

Because the primary purpose of this study was to evaluate backwashing
techniques, careful notes were recorded during the backwashing of
each filter after every run.  Bed expansion, backwash rate, air-scour
rate (where applicable), surface washer line pressure, water tempera-
ture, and media heights before and after backwashing were noted for
each filter.  Comments concerning the condition of the bed before,
during, and after cleaning were also recorded.  The comments were
particularly important because they contained information with respect
to surface cake formation, cracks in the media, mud ball or agglom-
erate formations, and other unusual qualitative observations.

Maintenance.  Periodic maintenance was required in various parts of
the pilot plant to ensure proper operation and collection of meaning-
ful data.  Because the sewage being filtered was a rich source of
nutrients and the pilot plant was constantly exposed to the sun,
algal growths flourished if not properly controlled.  To prevent
algal growths from occurring in piezometer tubes and composite sam-
pler lines, a 10 to 20-ml dose of 57, NaOCl (household bleach) was
introduced to each line on a once weekly basis.  After allowing a
15 to 20 min contact  time, the lines were drained and thoroughly
flushed.

Because the filter housings had plexiglass fronts, some means of con-
trolling algal growths within the filter housing was necessary.  To
retard algal accumulations, therefore, removable 1/4-in. plywood
covers were constructed to fit over the glass front of each filter
housing and prevent the passage of sunlight.

Although the covers were effective in preventing algal growths within
the filters, they could not prevent the accumulation of biological
solids on the inside  face of the plexiglass fronts.  If left unat-
tended, such growths  became thick enough to prevent one from viewing
the media.  It was necessary to scrape the growths from the glass at
least once a week using a rubber squeegee lowered into the  6-in.
handhole in the top of each filter housing.

A particularly  troublesome and persistent problem encountered was the
clogging of the piezometer taps.  This was probably caused  by  an ac-
cumulation of biological  solids on the fine  steel mesh  soldered over
                                  109

-------
 the  internal end of each 1/4-in. copper  tube.  It was found that one
 could generally free the tap by fitting  a  filled water bottle, with
 stem removed, over the end of the tap and  applying hand pressure.  If
 this was unsuccessful, a length of 1/4-in. tubing was fitted to the
 tap  and connected to the well hydrant.

 Biological solids coated the surface of  the influent rotameter tube
 and  float.  Solids accumulating on the float increased its drag,
 causing it to rise higher in the tube and  give erroneously high read-
 ings.  Daily cleaning was required to enable the meter to function
 properly, and this proved particularly inconvenient because of the
 construction of the rotameter.  Therefore, at the end of Phase II the
 rotameter was removed from the flow scheme, and flows were monitored
 only after the flow splitter, as shown in Fig. 23, for all remaining
 phases of the study.

 Results_^ Phase I

 Various parameters were used to compare  the backwash effectiveness
 between the two filters.  The most direct measurement was the abra-
 sion test, which revealed the relative amount of solids left on the
 media after a backwash.  Visual observations of the filter media
 through the transparent filter wall also directly indicated the ef-
 fectiveness of the backwash.  Indirect measures included the initial
 head loss of the filter media after backwash, the head loss develop-
 ment patterns and the filtrate quality.  Excessively dirty filter
 media or the presence of agglomerates and mud balls should result in
 higher initial head losses, more rapid head loss development, shorter
 filter runs and potentially poorer filtrate quality.  In addition,
 dirty filter media will alter the backwash rate required to achieve
 a particular degree of expansion.  Coatings should reduce the effec-
 tive density of the grains and reduce the  required wash rate.  Ag-
 glomerates will cause erratic behavior because portions of the bed
 may  not fluidize above the agglomerates, and because part of the
 media is held in the agglomerate and is not free to be fluidized.

 The  observations related to these measures of backwash effectiveness
 are  presented in the following paragraphs.

 Visual observations.  Small mud balls that appeared to be mixtures of
 sand, coal, and alum floe began forming during the first few filter
 runs.  The mud balls started out on the surface of the coal.  By the
 end  of the sixth filter run and backwash,  the mud balls had grown
 from 1/16 in. in diameter until 1/4 to 3/4-in. balls covered over 907»
 of the filter bed surface.  Before run 10  some of the mud balls were
 up to 1-1/2 in. in diameter and had sunk into the coal layer.  By
 the  end of run 14 almost all of the mud balls had sunk into the coal
 and very few remained on the surface.  A layer of fine sand had
worked its way to the surface of the coal, presumably due to reduc-
 tion of hydraulic subsiding velocity of the sand by the coatings
which had developed on the sand grains.
                                  110

-------
Just before the air scour was used (run 27), both filters had large
numbers of mud balls up to 3 in. in diameter at the sand-coal inter-
face.  The first use of air scour dramatically reduced the number of
mud balls in the south filter, and the largest one left was 3/4 in.
in diameter instead of 3 in.  After the backwash for run 29, three
backwashes after the first air scour, only one mud ball was observed
in the south filter, while numerous mud balls up to 2 in. in diameter
remained in the north filter.  The layer of fine sand had also dis-
appeared from the surface of the south filter.

Starting with the backwash for run 51, a layer of fine sand was once
again noticed over the coal in the south filter.  Also, a portion of
the media near the plexiglass front did not easily fluidize in the
water wash.  However, only a few small mud balls were seen in the
coal from then until the end of Phase I.

The north filter developed a heavy layer of fine sand over the coal.
As the head loss increased during filter runs, large cracks up to
3/4-in. wide and several inches deep formed at the media surface.
While most of the cracks extended along the filter walls, some smaller
ones appeared toward the center of the media surface.  The south fil-
ter had developed some cracks at the media  surface, but they were
much smaller and occurred less frequently than those in the north
filter.  After the fine sand was skimmed from above the coal, the
cracks at the media surface stopped forming.  However, another layer
of fine sand soon worked its way to the top of the coal and the cracks
at the surface of the north filter reappeared.

The sand used in the filters was somewhat small and undoubtedly con-
tributed to the difficulties encountered in keeping the filter media
clean and to the migration of fine sand to  the surface.  The use of
a larger sand would have been better.  Also, it should be recalled
that for this experimental Phase I, the sand was not skimmed to re-
move fines, which may have contributed to the sand migration problem.

Abrasion tests.  Figure 26 shows the amount of suspended solids re-
leased from the core samples of filter media as indicated by the
abrasion test results.  Two values are shown for the south filter at
run 27 because the normal water fluidization backwash was followed
by an air-scour assisted backwash as previously described.  A core
sample was taken after each backwash.  The  abrasion test results
started out at about 4 mg/g and increased to 22 mg/g.  The use of
air scour greatly reduced the abrasion test results in the south
filter where values dropped from over 11 mg/g at run 27 to less than
2 mg/g at run 34.  After run 45, the abrasion test results for the
south filter started to increase.  The lines shown in Fig. 26 are
least square fits using all of the values for the north filter and
the values from runs 45 through 72 for the  south filter.

The benefit of the air-scour auxiliary in cleaning the filter is
clearly evident from the abrasion test data.  However, the  gradual
                                  111

-------
    20
 |
 O
 CO
 O
 _j
 O
 co
    10
  Fig. 26.
              o NORTH FILTER - WATER FLUIDIZATION ONLY
                                                       ~
                SOUTH FILTER -AIR-SCOUR AUXILIARY  o
                                                       oo  o
             10     20      30     40      50
                              FILTER   RUNS
                                      60
                                           70
                     I
                    1
              i  rn   r r    i   n   i   i    i   r
             11132022261    10 18 25  1   7  15 20
              JUNE         JULY    AUGUST
MONTH AND DATE, 1973 (NONLINEAR SCALE)
  i
 26
MAY
80
Abrasion test when filtering  secondary effluent treated
with alum for phosphorus  reduction in Phase I.
deterioration evident  toward  the end of Phase I indicates that even
the air-scour auxiliary as used in this phase was not totally effec-
tive in keeping the media clean.

Initial head loss.   As mud balls form, the volume of void space in
the filter is reduced.  This  results in greater water velocities be-
tween the grains of filter media and therefore greater initial head
losses.  Therefore, an increase in initial head loss would indicate
the accumulation of mud balls in a filter.

At the beginning of Phase I when both filters were clean, the initial
head loss was 0.44  ft  for both filters.  Table 9 summarizes the means
and standard deviations of the initial head losses for various peri-
ods of filter operation in Phase I.  During the 3-1/2 weeks at the
beginning of the project, the values of initial head loss progressive-
ly increased.  In the  week immediately preceding the start of air
scour in the south  filter (filter runs 22 through 27), the average
initial head loss was  0.82 ft for the north filter and 0.80 ft for
the south filter.  The difference between the filters was negligible.
                                 112

-------
Table 9.  Initial filter head losses during various portions of the
          study (ft of water).

                           N filter, water      S filter, air-scour
 Runs                     fluidization only          auxiliary
22-27
28-45
46-57a
58-65
68-73
Mean
Std. dev.
Mean
Std. dev.
Mean
Std. dev.
Mean
Std. dev.
Mean
Std. dev.
0.82
0.096
0.64
0.141
0.55
0.057
0.72
0.144
0.54
0.062
0.80
0.105
0.44
0.036
0.37
0.061
0.52
0.114
0.41
0.029
n
 During this period the flow rate was only  1.6 gpm/sq ft instead of
 2 gpm/sq ft as for the other periods.

After  the south filter was backwashed using air  as  auxiliary agita-
tion,  the initial head loss on  the following filter run dropped
nearly to that observed at the  very beginning of the project.  In
the next 3-1/2 weeks, the initial head  loss for  runs 28 through 45
averaged 0.64 ft for the north  filter and 0.44 ft for the  south fil-
ter.   From  the eleventh week of operation through the twelfth week
of operation the initial head losses for both filters increased.
For these eight runs, 58 through 65, the north filter averaged 0.72
ft and the  south filter averaged 0.52 ft.   By this  time a  quantity
of fine sand had worked its way to the  top  of the coal layer.  It was
felt that this fine sand was adversely  affecting the head  loss char-
acteristics of the filters, therefore,  it was skimmed out  of the fil-
ters.  Not  surprisingly, the initial head losses for both  filters
decreased.  For the next six runs, 68 through 73, the north filter
averaged 0.54 ft and the south  filter averaged 0.41 ft of  initial
head loss.  The initial head losses in  the  filters  after skimming out
the fine sand should not necessarily,be the same as the initial head
losses at the beginning of the  experimental work because the media
characteristics were changed somewhat by the skimming operation.

The initial head loss data above supports the visual and abrasion
test observations.  The benefit of the  air-scour auxiliary compared
to water fluidization only is clearly demonstrated. However,  a
slight deterioration is evident in the  south filter, even  with the
air-scour auxiliary.

Head loss development.  Before  the use  of air scour was  started,  the
head loss in each filter was similar with respect to time  and  to
                                  113

-------
depth in the filter.  Figure 27 shows typical head loss versus time
curves for various intervals of media depth, measured from zero depth
at the top surface of the media.

After air scour was begun on the south filter, it developed head loss
at a slower rate than the north filter.  Figure 28 is a typical exam-
ple.  During the two weeks following the beginning of the air scour
(runs 28 through 44), the filters were generally backwashed before
they reached 4 ft of head loss.  Later they were operated to higher
head losses.  Starting with run 45, terminal head losses for both
filters were about equal, with the south filter frequently having a
slightly greater head loss.  This was particularly true when higher
head losses were reached.  Figures 29 and 30 are typical examples.
After run 65, 8-1/2 weeks after air scour was started, the layer of
fine sand that had worked its Way into the coal was removed from the
filters.  Figure 31 shows a typical curve of head loss versus time
after the fine sand was removed.

The effect of the air-scour auxiliary on head loss patterns is
clouded by the variable extent of surface cracks in the filters.  If
surface cracks were absent and mud balls were present, the dirtiest
filter should have the highest rate of head loss development.  A
higher rate of head loss development for the dirtier north filter was
not consistently observed in this Phase.  This anomaly is explained
as follows.

During the first four weeks after air scour was started, the filter
runs were generally terminated when the head loss reached only 3 or
4 ft.  It was then observed that, at greater head losses, the head
loss in the south filter would approach and even surpass that in the
north filter.  This was attributed to the fact that the north filter
experienced more extensive surface cracking during the filter runs.
These cracks allowed the surface to be bypassed and solids removal to
take place in deeper layers of the filter. The head losses shown in
Figures 29 and 30 clearly indicate that most of the removal in the
south filter took place in the top 4 in. of the media, while in the
north filter the removal took place over the upper 8 in.  This indi-
cates that the greater head loss in the south filter on some occa-
sions was not due to accumulated dirt, but was rather due to the fact
that the dirtier media in the north filter resulted in more extensive
surface crack formation.

Backwash rates required.  For runs 31 through 63, both filters were
backwashed to a constant expansion of 38 to 40% and the backwash rate
needed to achieve this was recorded.  For runs 31 through 49 the
north filter required an average of 19.4 gpm/sq ft, and the south
filter required an average of 22.2 gpm/sq ft.  For runs 51 through 61
the required backwash rates were 19.7 gpm/sq ft for the north filter
and 20.0 gpm/sq ft for the south filter.  Table 10 lists the means
and standard deviations of the backwash rates of the filters for the
                                 114

-------
         NORTH FILTER
         (WATER FLLUDIZATION ONLY) 0 to 24 in.
GO
«/»
B
a
&
X
                                     4 to 24 in
8 to 24 in.
                                    12 to 24 in.
                                    16 to 24 in.
                 I	I
            SOUTH FILTER
            (WATER FLUIDIZATION ONLYY
 0 to 24 in.
               I       I
                                                                              4 to 24 in.
     05     10     15     20
                    TIME, hrs
 25      0     5     10     15    20
                         TIME, hrs
25
              Fig. 27.  Head loss vs time at various media depths, run 27.

-------
         NORTH FILTER
     WATER FLUIDIZATION ONLY)
   3_
 4>
oo
to
Q
                                    0 to 24 in.
                   I
I
                                     4 to 24 in.
                                    8 to 24 in.
                                    12 to 24 in.
I
                           SOUTH FILTER
                          (AIR-SCOUR AUXILIARY)
                                                      0 to 24 in.
                                                      4 to 24 in.

                                                      8 to 24 in.
                                                    012 to 24 in.
I
                  10      15     20
                      TIME, hrs
              25
                             10     15      20
                                 TIME, hrs
      25
                Fig. 28.  Head loss vs time at various media depths, run 42.

-------
 g  4
in  o
«/>  o
O
      NORTH FILTER (WATER FLUIDIZATION ONLY)
                                     0 to 24 in.
                                     4 to 24 in.
                                     8 to 24 in.
                   I     I     I     I     I
 SOUTH FILTER  (AIR-SCOUR AUXILIARY)
                                    0 to 24 in.
                                                                                     4 to 24 i
                                                                                            in.
                                                                                     8 to 24 in.
                                       I
     05    10    15   20    25  30   35  40
                        TIME, hrs
0    5    10
15   20   25   30
     TIME, hrs
35   40
                  Fig. 29.  Head loss vs time at various  media depths, run 59.

-------
oo
                NORTH FILTER  (WATER FLUIDIZATION ONLY)  SOUTH FILTER  (AIR-SCOUR AUXILIARY)

                                                                                            .0 to 24 in.


                                               ,0 to 24 in.
                                                                                            4 to 24 in.
                                                                                            8 to 24 in.
                                                                                            12 to 24 in.
                             10     15      20
                                  TIME,  hrs
25
10      15
   TIME, hrs.
20
25
                          Fig. 30.  Head loss vs time at various media depths, run 63.

-------
 4)

1
t/J
uo
2
Q
       NORTH FILTER  (WATER FLUIDIZATION ONLY)   , SOUTH FILTER
                                                      (AIR-SCOUR AUXILIARY)

                                           0 to 24 in.
    5-
4 to 24 in.
                                           8 to 24 in.

                                          12 to 24 in.
                                                to 24 in.
                                                                                        4 to 24 in.
                                                                                         8 to 24 in.
                                                                                        12 to 24 in.
      0
10   15   20   25   30   35   40     0
           TIME,  hrs
                   10    15   20   25   30
                            TIME, hrs
35   40
                   Fig.  31.   Head loss  vs  time at various  media depths,  run 71.

-------
 Table  10.  Backwash rates required to achieve 38 to 40% bed
            expansion  (gpm/sq ft).


                           N filter, water       S filter, air-scour
  Runs                    fluidization only           auxiliary
31-49
50-61
Mean
Std. dev.
Mean
Std. dev.
19.4
0.88
19.7
1.85
22.2
0.80
20.0
1.07
 two periods.  The backwash temperature averaged 22.9 °C for runs 31
 through 49 and 23.1 °C  for runs 50  through 61.

 The data presented support the other observations concerning  the
 cleanliness of the filter media.   Shortly after initiation of air
 scour, the media of the south filter was cleaner and thus had a
 higher average density than  the north filter.   The higher density  re-
 quired a higher backwash rate to  achieve a given expansion.   Later,
 as coatings developed on the air-scoured media,  the average density
 decreased and the required backwash rate decreased. Thus, the re-
 sults of Table 10 are in harmony  with the abrasion test results and
 initial head loss results.   All three tests  demonstrate the superi-
 ority of air-scour auxiliary over water fluidization alone, but all
 three also indicate some deterioration of the  air-scoured filter
 towards the end of Phase I.

 Water quality.   The means and standard deviations of the  various
 water quality parameters are given  in Tables 11  and 12.   Results for
 each  parameter were divided  into  two periods.  The first  period
 (Table 11)  covers the initial 3-1/2 weeks of filter operation when
 both  filters  were backwashed by water fluidization only.   The second
 period (Table 12)  covers the remaining 8-1/2 weeks of filter  opera-
 tion  when the south filter received air-scour  auxiliary.   Except for
 turbidity,  the  values for the solids contact unit influent may not
 be used  to  calculate  actual  removals through the  treatment process.
 Samples  of  influent to  the solids contact unit were grab  samples,
while  the other  samples  were  composited over approximately one  day.
All of the  turbidity measurements were made  on grab  samples taken  at
 approximately  the  same  times  for each point  in the treatment  process.

From  the data on filter  effluent quality, no apparent differences
are evident between the  two  filters.  One would expect the dirtier
filter to yield poorer filtrate due  to surface cracks permitting
deeper penetration of solids, and due to higher interstitial veloci-
ties caused by mud balls and agglomerates.  The absence of detriment
to filtrate quality in this work must be attributed to the fine grain
size of the media, the low filtration rates  and the low terminal head
                                 120

-------
Table  11.   Results of analyses  during alum treatment  (Phase  I)  for  samples  from May  17  to July  11,
            1973, when both filters  were washed  by water  fluidization only.   (All  results from com-
            posite samples  except solids contact influent.)
Solids contact
influent
Avg suspended solids (mg/1)
CTa (N - 16)
Avg turbidity (FTU)b
aa(N = 55)
Avg BOD5C (mg/1)
<7a(N = 13)
Avg TOG (mg/1)
aa(N = 15)
Avg total P04 (mg/1)
aa(N = 8)
Avg ortho P04 (mg/1)
'
-------
r-o
       Table 12.   Results of analyses during alum treatment series (Phase I) for samples from July 11 to
                  August 20, 1973, when air scour was being used on the south filter.   (All results from com-
                  posite samples except solids contact influent.)
Solids contact
influent
Avg suspended solids (mg/1)
aa (N = 4!)
Avg turbidity (FTU)b
a (N = 207)
Avg BOD5C (mg/1)
CT (N = 14)
Avg TOG (mg/1)

-------
losses.  Had one or more of these filtration variables been increased,
the detriment may have occurred.

Statistical comparisons to determine differences in the effect of
backwashing on filter performance efficiency betveen the filters are
not reported since a proper statistical base did not exist in this
experimental study.  Since the filter media was not returned to
exactly the same condition following each backwash, the individual
runs on a filter cannot be considered independent, which is a funda-
mental assumption in statistical theory.  Because it was desired to
determine the cumulative effects of the various backwashing methods
over a period of continuous operation of the filters, it was not
practical to thoroughly restore the media after each filtet run, and
it was equally impractical to replace it.  The minimum effort alter-
native for the application of statistical comparisons would be to
run at least two separate eight-week studies using identical media,
switching the backwashing techniques on each housing, and replacing
or completely restoring the media at the end of each eight-week
series.  Time and expense considerations precluded the additional
study, so no base exists for making statistical inferences.

Clean up operations at the end of phase I.  The results of the sus-
pended solids concentrations versus quantity of backwash water for
the first backwash following the use of air scour  (run 27) are shown
in Fig. 32.  Since a normal backwash using only water fluidization
immediately preceded this backwash, the area under the curve repre-
sents the additional suspended solids released by  the first air scour.
The suspended solids removed by the air scour was  109.6 g per sq ft
of filter area.  If the suspended solids are assumed to come from a
12-in. layer of coal and a 1-in. thick layer of sand, 4.13 mg sus-
pended solids were removed per gram of filter media.

The amounts of suspended solids released from the  filter media in
the final cleanup operations after run 78 are shown in Table 13.
Based on the same amount of filter media as was assumed before, 13.3
and 3.2 mg of suspended solids were released per  gram of media for
the north and south filters respectively.  All mud balls were broken
up during the cleanup operation.  The series of cleanup steps was
continued until the media appeared to be in new condition and the
last  step appeared to release very few  additional solids.  The sub-
stantial difference in total suspended  solids released clearly shows
that  the south filter which had routinely been washed with air-scour
auxiliary was in much cleaner condition  than the  north filter.  It
should be borne in mind that this final  cleanup operation came after
a week of filtering uncoagulated trickling filter effluent at the end
of the alum treatment series.

Summary and Conclusions - Phase I

The objectives of  this phase were to evaluate  the effectiveness  of
filter cleaning by water fluidization backwash  alone  as compared  to
                                  123

-------
         CO
             loooh
              800h
              60°
10
        co
        9
        _i
        O
        CO
         Q
         z
         UJ
         a.
         CO
         :D
         CO
              400K
              200h
0      10


   Fig. 32.
                                20      30      40      50     60      70      80
                                    VOLUME OF BACKWASH WATER USED, gal./sq ft
90
100
                              Suspended  solids concentration of backwash water vs quantity
                              of backwash water used, run 27, second backwash of the south
                              filter immediately following the first application of air
                              scour.

-------
Table  13.   Suspended solids released from filters in special back-
            washes after run 78.
                                          g of SS removed/sq ft
                                              of  filter area
      Backwash used
North filter
South filter
Air followed by water
Air and water combination
Air followed by water
Air and water combination
Air and water combination
Total
87
144
57.5
47
28. 2b
393.7
37.3
28.2
a
14.4
8.5
88.4
  Step omitted.

  Backwash conducted but no suspended solids value obtained therefore,
  the value indicated is based on  the comparison of turbidity with
  the other samples.

air scour followed by water fluidization backwash, and to compare the
performance and operation of filters backwashed by these methods when
filtering wastewater.  For this study two dual-media filters were
used following alum treatment of domestic wastewater for phosphate
removal.  One filter used air scour followed by water backwash, while
the other used water fluidization backwash alone.

The media in the filter backwashed with water only became very dirty,
while the media in the filter with air scour during backwash was much
cleaner.  However, toward the end of the project, the media in the
air-scoured filter also showed some deterioration.  Few mud balls
were observed in the air-scoured filter while numerous, large mud
balls were observed in the other filter.  When operated at only 3 to
4 ft of head loss the filter backwashed with water only had greater
head losses than the filter with air scour.  When operated to greater
head losses this was not true due to the formation of surface cracks
in the filter receiving water backwash alone.  Little difference in
effluent quality was observed between the two filters.

The study of the filtration of domestic wastewater which had been
subjected to secondary treatment and subsequent alum treatment for
phosphate reduction leads to the following conclusions concerning
backwashing:

1.   Filter media were kept cleaner by air scour than by water fluid-
     ization backwash alone as evidenced by the results of visual
                                  125

-------
     observations, media abrasion tests, initial head losses, back-
     wash flow requirements, and the final cleanup operation.

2.   Some minor deterioration of the condition of the media of the
     air-scoured filter was evident toward the end of this experi-
     mental phase as evidenced by the results of the media abrasion
     test, initial head losses and backwash flow requirements.

3.   Air scour prevented the formation of mud balls, while water
     backwash alone did not.

4.   The two methods of cleaning resulted in little difference in
     effluent quality.  The use of a coarser filter media, a higher
     filtration rate, or higher terminal head loss may have demon-
     strated that a cleaner media would give better filtrate quality.

5.   The sand used in the experimental filters was smaller than de-
     sirable (0.38-mm effective size) and was not skimmed of fines
     on placement.  This contributed to backwashing difficulty and
     sand migration to the surface of the coal of the dual-media
     filters.

                   Operation and Results - Phase II

             Dual-Media Filtration  of  Secondary Effluent

Phase II was a comparison of three methods of backwashing during the
direct filtration of secondary effluent at the Ames, Iowa, trickling
filter plant.  All three filters were equipped with dual media.  One
was washed by water fluidization only.  The second was washed by air
scour followed by water fluidization, hereinafter referred to as air-
scour auxiliary.  The third was washed with a rotary surface wash
auxiliary which operated before and during the water fluidization
backwash.

Prior to commencing Phase II, the west filter with surface wash
auxiliary was installed and new filter media was installed in all
three filters.   This was done to ensure that Phase II would not be
influenced by any carry-over effects from Phase I.

Operation - Phase II

The secondary effluent of the Ames trickling filter plant was pumped
directly to the influent splitter box and thus became the filter in-
fluent as illustrated in Fig. 23.

Filters.  Because it was desired to observe the effects of continuous
operation on filter efficiency, the pilot plant was operated without
major interruptions from August 28 to October 30,  1973.  The entire
series of runs  was designated Phase II, with individual runs being
numbered consecutively from 1 to 64.  Normal procedure dictated one
                                 126

-------
observation run per week at which time the head loss development,
influent and effluent turbidities, and flow temperature were moni-
tored, and composite samples were collected.  During the remaining
runs of a given week, only initial readings immediately following
backwashing were recorded.  The backwashing observations outlined
in Phase I were recorded for each backwash.

When the series was begun it was hoped to be able to operate the
filters at a loading rate of 2 gpm/sq ft.  Piping losses and solids
accumulations in the lines gradually reduced this figure as operating
time increased to 1.6 gpm/sq ft near the end of Phase II.  Even
though the rate declined gradually, the flow split to the three fil-
ters was equal, so that valid comparisons between the filters are
possible.

Backwashing.  Ordinarily one would backwash filters when the head
loss had reached some maximum permissible value or on a time cycle
prior to attainment of the maximum permissible head loss.  Neither of
these procedures was deemed practical during Phase II because the
runs were quite short and too much operational time would be re-
quired.  Typically, the filters required approximately 10 to 12 hr to
attain maximum head loss (point of incipient bypass in the splitter
box).  Furthermore, divergence in run lengths was expected to become
more pronounced as operation time increased.  Therefore, a schedule
was adopted whereby the filters were backwashed and placed in service
each morning as a group.  This meant that by late evening the filters
had begun to bypass through the lower half of the splitter box and
that by the next morning nearly all of the influent was being by-
passed.

Although each of the filters was backwashed using a different tech-
nique, all three filters required similar initial preparations as
described previously in the Phase I operational discussion.

A nominal water fluidization backwash rate of 20 gpm/sq ft was se-
lected for use on all three filters, this value being an upper limit
for normal backwashing of filters used in water treatment.  The rate
was easily obtained during the first 20 runs, but after that the
attainable rate decreased, presumably due to solids accumulations in
the backwash lines and partially clogged underdrain strainers.  From
run 21 on the rates varied from 18.5 to 20 gpm/sq ft, with a pre-
ponderance of values in the 19.0 to 19.5 gpm/sq ft range.  Attempts
to clean the distribution nozzles were not made because to do so
would have required all media to be removed from the filter housing.

Backwashing of the north filter consisted of a water fluidization
(only) backwash at the 20 gpm/sq ft rate for a duration of 5 min and
was not preceded by air scour or surface wash.  Bed expansion
during fluidization varied from a low of 8.3% to a high of 37.5%,
with an average value-of 25.6%.  The slightly reduced backwash rates
had no noticeable influence on bed expansion, which appeared to be
                                  127

-------
more a function of bed condition, i.e., number of mud balls and
agglomerates.

The south filter was subjected to air scouring at a rate of 3.72
scfm/sq ft for 5 min prior to the water (only) backwash.  Before the
air line valve to the filter was opened, the filter was drained down
through the effluent line until the water level was 6 in. below the
backwash water collection trough.  This prevented the loss of media
out the backwash waste line.  By run 36 this procedure had proved
unsatisfactory because a surface coating formed on the anthracite
which made draining the water above the filter nearly impossible.
More importantly, the filter itself was draining below the surface
cake, causing a negative head condition in the interior of the media.
This caused air to be drawn in through the piezometer tubes, which
air bound the filter and impaired the air-scour agitation.  To combat
these difficulties in subsequent backwashes, the procedure was al-
tered by first "bumping" the bed with a short, low volume application
of air.  This practice broke the surface layer of the media suffi-
ciently to allow proper draining without negative head development,
after which the air scour could proceed in the normal fashion.

Upon completion of the air scour, the air line valve was closed and
the backwash line valve opened to a rate of 20 gpm/sq ft for a total
duration of 5 min.  Expansion for the south filter ranged from 25.5
to 42.6%, with an average value of 34.0%.

The west filter backwashing procedure was initiated with a 2-min solo
operation of the rotating surface washer.  Next the 20 gpm/sq ft
water backwash rate was applied to the filter and was operated in
combination with the surface washer for a total of 3 min.  Finally,
the surface washer was shut off and the water (only) backwash was
continued at the initial rate for an additional 2 min so that both
the surface wash and water backwash were operated for 5 min each.
Expansion of the west filter during the water fluidization (only)
phase ranged from 25% to 50%, with an average value of 37.4%.

Sampling procedures and data collection.  The frequency of observa-
tion runs in Phase II was less than in Phase I to reduce budget ex-
penditures.  Observation  runs in Phase II were conducted once weekly
and consisted of a careful monitoring of performance parameters
throughout the duration of the run.  Flow rates, influent and efflu-
ent turbidities, flow temperatures, and head loss development were
monitored at 1 to 2-hr intervals.  Prior to the beginning of each ob-
servation run the normal maintenance, described in a later section,
was performed to ensure the collection of meaningful data.  The auto-
matic composite samplers were also turned on and checked before
placing the discharge lines into refrigerated sample receivers.  At
the conclusion of each observation run, core samples were collected
for the abrasion tests and influent and effluent composite samples
collected for laboratory analyses.
                                 128

-------
In all cases the filter effluent samples were composited for the du-
ration of the filter observation run.  However, for the first six ob-
servational runs (2, 14, 22, 29, 36, 43) the composite samplers were
left on for the entire 24-hr period even though the filters began to
bypass after about 12 hr.  Because it was thought that this procedure
may have generated misleading data, the routine was changed for the
last two observation  runs so that each composite sampler was shut
off when the filter first started to bypass.

Filter influent samples were not composited for the majority of the
runs because filter performance was not considered a prime objective
of the research, and because adapting the available sampler to the
pilot plant posed some difficulties which were not solved until later
in Phase II.  A grab sample was taken at the conclusion of each ob-
servation run for numbers 14, 22, 29, 36, 43.  Admittedly, this was a
weak procedure, even for a parameter of secondary importance, but the
influent data were, fortunately, later obtained from the log sheets of
the Ames Water Pollution Control Plant (composite samples).  For the
final two observational runs the filter influent was automatically
composited, and the sampler was shut off after the last filter began
to bypass.  The influent was also composited during the first obser-
vation run  (run 2), which occurred while the influent was still being
pumped through the erdlator of the mobile water purification unit used,
in Phase I.  This practice was discontinued after run 3 of Phase II.
Although the influent was composited over the entire 24-hr period
during observation run 2, no unusual differences were noted between
the sample  analyses of this run and the last two runs.

Other details of sampling, data collection, and maintenance were
identical with Phase I.

Results - Phase II

Visual observations.  Although the condition of the media varied sub-
stantially  among the three filters, certain characteristics were
common to all three.  One common trait, the accumulation of mud balls
and agglomerates, was a problem encountered in varying degrees
throughout  the study.  Because of the downflow configuration of the
filters and the relatively small particle size of the anthracite, the
majority of the suspended solids removal took place in the uppermost
4 to 6 in.  on the anthracite layer.  As each run progressed, the
solids in the influent filled the interstices of this upper layer and
simultaneously compressed the layer into a tightly packed mat as the
head loss across the layer increased.

Because of  the biological nature of the influent solids and the ad-
sorptive properties of the anthracite, the particles of media in the
matted layer were bound tenaciously together, forming a plug at the
surface which required violent agitation for its complete disintegra-
tion.  Since this was not always accomplished by a given backwash
technique or within a given backwash cycle, the matted layer was more
                                  129

-------
 often  split  into  fragments which descended  to  the  sand-coal  interface
 during fluidization  and were  then designated as mud balls  as  they
 floated at the  interface  of the fluidized bed.  Those  fragments which
 affixed themselves to  the sides of  the  filter housing  were designated
 as  agglomerates.  Nomenclature aside, all fragments observed  during
 fluidization were spawned from the  matted layer as a direct result of
 the overall  ineffectiveness of the  backwash technique  being used.

 When present in sufficient concentrations,  the mud balls and  agglom-
 erates caused similar  types of problems, in varying degrees of
 severity, in all  three filters.  If not broken up by the preceding
 backwash cycle, these  accumulations shortened filter runs by  increas-
 ing the initial head loss and reducing  interstitial volume of the bed
 available for subsequent  solids removal.  During the backwash cycles,
 the mud balls and agglomerates prevented uniform distribution of water
 or  air and forced them to form high velocity jets or streams.  Be-
 cause  of this channeling  effect, areas  of poor fluidization developed
 above  the fragments which reduced the effectiveness of the wash and
 created conditions favorable for the formation of additional  mud
 balls.

 Not once in  the course of this study were mud balls observed  to have
 originated in the sand layer of any of  the  filters.  All were com-
 posed  originally of anthracite and  solids trapped during the  course
 of  the filter run, and were formed  as previously described.   Sand was
 observed in  some of the mud balls and agglomerates, but this  was
 caused by the jets of water which tended to lift the sand high into
 the coal layer  where it became trapped.

 Surface  layers  of solids  on the very top of the media  were observed
 frequently on all the filters just  prior to backwashing.

 Surface  cracking was observed in all three  filters but was most pro-
 nounced  and  sustained in  the north  filter.  Cracks in  the north fil-
 ter were generally 1/4-in. wide, 3  to 7-in. long, and  usually appeared
 to result from  the media  pulling away from  the filter  walls as the
 bed compressed.  More severe cracking in the north filter was observed
 on three occasions when the cracks  were from 1/4 to 1/2-in. wide and
 extended the full 18-in. width of the filter.  Surface cracks in the
 south  filter were less severe and less  frequent than those in the
 north  filter.   Slight cracking, usually 1/16-in. wide  and 1 to 2-in.
 long, was noticed, but only about one-third as often as in the north
 filter.  The west filter was essentially free of surface cracks
 throughout the  study.

Dead spaces were observed at the bottom center of the  sand layer in
 each filter.   The sand in this location, between two outside  adjacent
underdrain distribution strainers,  did not fluidize during the back-
wash.  No difficulties were encountered as a result of this.
                                  130

-------
Visual observations - north filter.  Because of the relative inef-
fectiveness of the water fluidization (only) backwash, the accumula-
tion of mud balls and agglomerates in the north filter was rapid and
sustained.  During the backwashing cycle of the first five filter
runs, the mud balls observed ranged in size from about 2 to 6 in. in
diameter and appeared in number of about 10 to 20.  By the end of the
fifteenth run, however, the size and concentration of the mud balls
and agglomerates had increased to the extent that fluidization and
stratification during backwashing did little more than shuffle and
shift the mud balls slightly.  Actual cleaning of the media was
virtually halted.  From run 16 until the completion of the study,
mud balls and agglomerates consistently composed 50 to 7070 of the
volume of the anthracite layer visible at the window of the filter.
Typical comments as recorded in the data book are presented below for
runs 17 and 50, respectively.

            Heavy surface coating.  No cracking.  4-in.
            penetration into bed.  Large mud ball at
            start 18 in. x 8 in. (almost entire coal
            layer).  Smaller mudballing also present.
            Bed poorly stratified.

            1/4-in. cracks, length of glass, heavy sur-
            face coating.  Intermixing and agglomerates
            70-80% of bed.  Fluidization poor, huge
            agglomerates - 24 in. x 6 in. x 8 in., and
            6-12 in. diameter - causing jets and inter-
            mixing.  Large agglomerate is only 3 in.
            from bottom of filter.  Bed essentially
            intermixed - agglomerates did not break.

Occasionally the concentration of mud balls and agglomerates was even
higher although this higher level (80 to 9070) was not sustained for
any length of time.  The complete clogging of the media was probably
prevented in part by anaerobic decomposition within the large masses
in the filter.  The distinct odor of hydrogen sulfide was detected
frequently while backwashing the filter and lent some qualitative
support to this possibility.

On 10 occasions (runs 14S 17, 21, 39, 49, 51, 53, 56, 60, and 62), the
weakness of water (only) backwashing technique as applied to sewage
filters was dramatically demonstrated.  When the backwash water was
applied to the media, the entire matted layer, a block 18-in. square
and 4 to 6-in. deep, rose as a plug in the filter housing.  Three
times (run 14, 49, 62) the water backwash failed to break the layer
(even slightly), and it settled as one large mass to the sand-coal
interface.  In the remaining instances the matted layer was broken
into large chunks which formed mud balls or agglomerates.

Although the high concentration of mud balls and agglomerates in the
north filter had no apparent effect upon the quality of the filtrate,
                                 131

-------
 it  did  cause  the most  severe  instances  of high  initial head  loss  and
 shortened  runs.  Initial head loss  readings  averaged nearly  1  ft
 greater than  those  observed for  the other two filters, and filter
 runs were  noticeably shorter  as  will be shown later.  The large,
 solid masses  often  covered one-half to  two-thirds of the bed width
 and caused the backwash water to channel and to lift sand high into
 the anthracite layer where it became trapped.   Dead areas occurred
 above the  mud ball  layers and the unequal distribution of backwash
 water left the media mounded  and piled  after the backwash.   Fluidiza-
 tion of free media  was poor,  and stratification of the coal  and sand
 layers  became virtually non-attainable.  Frequently, large gaps and
 holes were left around the mud balls and agglomerates after  the wash
 which failed  to fill in with  filter media.

 Visual  observations -  south filter.  The media  of this filter  under-
 went a  series of significant  changes throughout the course of  this
 investigation.  Through run 5 the build up of mud balls and  agglom-
 erates  in  the bed had been minimal, roughly  a half dozen mud balls
 were observed during each backwash  ranging in size from 1 to 2 in. in
 diameter.   During this initial period the 5-min air scour appeared to
 be  thoroughly breaking up the matted layer and  intermixing the sand
 and coal layers.  Air-scour agitation was most  pronounced during  the
 first minute, during which time  layers  intermixed.  The bed  quickly
 subsided to a steady, pulsing action, primarily at the surface, for
 the remainder of the scour.

 Starting with run 6, however,  much  larger mud balls and agglomerates
 began to appear during the backwash.  While  the bed was being  fluid-
 ized these large chunks settled  to  the  sand-coal interface and chan-
 neled the  backwash  water to the  sides of the housing.  As it had  in
 the north  filter, the channeling action caused  sand to be carried and
 trapped  in the upper anthracite  layer.   Additionally, the effective-
 ness of  the air scour had decreased markedly, and the air appeared
 to  agitate only the top 1 to  2 in.  of the anthracite.  The air-scour
 agitation  was carefully observed  during the  backwash following run
 10.  The air was channeled readily  through the  media instead of
 being uniformly dispersed, and the  bed  as a  whole exhibited  a  gelati-
 nous or  cohesive character.   By  the end of the  air scour the agita-
 tion had usually begun to be  most effective, eroding a small portion
 of  the bed  and piling it on the  surface.  Even  more perplexing, the
media fluidized and stratified fairly well following the seemingly
 ineffective scour.

 Compared to results obtained  from Phase  I on alum-treated, secondary
 effluents,  the air-scour agitation  results were quite poor.  Initially
 this was attributed to the much higher  solids loading on the filter,
but a slight procedural change in the backwash  cycle of run  36 proved
otherwise.  Until this particular run,  the procedure had been  to
 lower the water level 6 in. below the washwater trough to prevent
media loss.  Because of the highly  clogged nature of the upper anthra-
 cite layer, this was a relatively slow  process,  so the procedure  was
                                 132

-------
altered for run 37.  During this backwash cycle the water level was
not lowered below the overflow weir, and the air was applied at the
standard rate.  The results were surprising because the air scour
suddenly regained its past effectiveness.  The matted surface layer
was effectively broken up as was a very large 4 by 18-in. agglomerate
at the interface.  Furthermore, the sand and coal thoroughly inter-
mixed during the air scour and then stratified excellently during
fluidization of the bed.

A subsequent investigation revealed the reason for the poor air
agitation during the previous 26 backwashes of the south filter.
Both the surface coal and the matted anthracite layer severely re-
stricted the ability of the bed to drain the water above, as evi-
denced by the long period required to do so.  This caused the water
within the media below the matted layer to drain faster than the
water above could pass through the mat.  Therefore, a negative head
condition occurred in the filter which, in turn, drew air into the
bed through the piezometer taps.  The entrained air tended to bind
the filter media, causing it to appear gelatinous and to resist
break up by the air agitation.

A simple procedural alteration was incorporated in the south filter
backwashing sequence for subsequent runs.  Before draining the media,
the bed was subjected to a short, low volume application of air to
break up the compressed surface layer.  This allowed the filter to
drain freely without inducing negative head and air binding of the
media.

Immediately following the procedural change, the condition of the bed
improved with respect to mud ball and agglomerate concentration.
However, smaller 2 to 3-in. diameter mud balls began to accumulate
and cause backwash water distribution problems during runs 39 and 40,
although these were quickly dispersed and reduced in number by run
41.  The next 20 runs were characterized by relatively few mud balls,
but were hampered by an inability of the air agitation to completely
disintegrate the upper clogged layer in the anthracite.  Typically,
the air scour broke the left and right one-third portions of the mat,
but left the middle fragment which settled to the sand-coal inter-
face.  This may be the result of poor air distribution by the filter
underdrain systems with strainers on about 7-in. centers.  These
fragments varied in length from 6 to 12 in. and in depth from 2 to
3 in., and interfered with proper fluidization and stratification of
the media.  Usually the agglomerate was broken up during the next
backwash cycle, but was immediately replaced by another.  The back-
washing entry for run 54 typified this period of operation.

            Minor surface cracking present.  Air scour
            does not break up mat.  At media interface
            mud balls sinking into sand.  Approx. 1/3
            of bed is not washing.  Bed frees during
            wash and media is well stratified.
                                  133

-------
For the remaining four runs, the above trend disappeared, and the air
scour appeared very effective in breaking the mat.  Several small mud
balls were observed but caused no serious problems with backwash dis-
tribution, intermixing, or fluidization and stratification.

Visual observations - west filter.  Except for the first two runs of
the study, the west filter was equipped with a rotating surface
washer as described previously in detail.  For the first two runs the
bed was subjected to surface wash from a fixed nozzle washer.  For
the next 29 consecutive runs a rotating jet washer was used which
consisted of an inverted, three-armed lawn sprinkler.  Finally, a
two-armed, two-nozzled, rotary washer was installed using nozzle
jets similar to those in actual surface washers.

Backwashes following runs 1 and 2 were characterized by one or two
large agglomerates (6 by 5 in. and 12 by 3 in.) accompanied by a host
of smaller 2 to 3-in. diameter mud balls.  The fixed nozzle washer
was breaking the mat into chunks which remained fairly well intact
for the rest of the wash.  When first used following run 3, the make-
shift surface washer appeared to do an excellent job in breaking up
the matted layer leaving only three, 2 to 3-in. diameter mud balls
visible in the bed.  The very next backwash revealed a highly plugged
surface layer which effectively resisted break up by the surface
washer.  When the bed was fluidized, the matted anthracite layer rose
as a plug with cross section equal to that of the housing and a depth
of about 8 in.  Below this plug approximately five 2-in. diameter mud
balls floated at the interface of the sand and coal layers.  Even-
tually the plug broke up into large agglomerates which sank to the
interface.

The condition of the bed stabilized during the next 10 backwashes to
one of several small mud balls, 3 in. in diameter or less, plus one
or two large agglomerates at the interface, sized approximately 8 by
2 in. to 6 by 2 in.  After a brief period of almost no mud balls or
agglomerates, this pattern was consistently observed from run 17 to
run 32.  Although these accumulations caused some channeling of back-
wash water and intermixing of media, the media continued, generally,
to fluidize and stratify quite well, as evidenced by this entry for
run 30:

            7-in. penetration into bed.  Heavy surface
            coating.  Mud balls in bed larger than in
            south filter.  4-in. dia., 2-in. dia., three
            3-in. dia., 8 in. x 2 in.  Bed is generally
            well stratified.

Preceding the backwash for run 32, the third surface washer was in-
stalled.  Although the action of the newly installed washer did not
appear particularly violent, the condition of the bed continued to
improve for the next three runs.  Then the familiar pattern of several
large agglomerates accompanied by approximately six small mud balls
                                 134

-------
returned with the backwash following run 36.  Except for a brief
period from run 46 to 49, during which the bed was relatively free of
mud balls and agglomerates, this trend continued for the balance of
the phase and is exemplified by the entry for run 43:

            No visible cracks, cannot see surface coating
            (blocked by support).  No appreciable inter-
            mixing.  Bed is well stratified.  One 5 in.
            x 1 in. agglomerate visible at sand-coal in-
            terface.  Surface washer working, violent
            swirling in area of washer when bed fluid-
            ized.  Several large mud balls observed:
            3 in. x 12 in. x 12 in., 6 in. x 6 in. x 3 in.,
            (2) 2-in. dia., (2) 1-in. dia.

Regardless of the washer used, therefore, the condition of the bed
remained relatively constant throughout the study - not good but not
exceptionally poor as in the north filter.  Several periods of signi-
ficantly improved condition of the media were observed, but were also
short-lived.  Fluidization and stratification of the media were not
severely hindered except when the accumulations became heavily con-
centrated, which was also infrequent.  Surprisingly, the action of
the surface washer was most effective after the bed had been fluidized
and the rotating arm was turning through the fragments of the mat,
and not when it was attempting to break up the surface of the static
layer.

Abrasion tests.  The abrasion tests were considered of prime impor-
tance in evaluating the effectiveness of the various backwashing
techniques because they directly indicate the condition of the filter
media.  As previously described, the abrasion test contained a step
in its procedure in which the core sample was subjected to a vigorous
30-min mechanical mixing.  Obviously, the mixing action tended to
break up some of the anthracite sample, so a background level was
determined for the test by subjecting an unused sample of anthracite
(out-of-bag) to the procedure.  The test was run twice on the same
sample of clean anthracite and yielded values of 0.61 mg/g and 0.22
mg/g, respectively.  The latter valve was accepted as the background
level because the former was influenced by coal dust which initially
adhered to the media.

The results of the abrasion tests are displayed graphically in Fig.
33.  Unquestionably, the most striking trend is that exhibited by the
north filter.  Except for a slight dip after run 20, the abrasion
test values for this filter climbed steadily throughout the filter
run series.  The rapid and sustained build up of solids on the north
filter media, as evidenced by these tests, is a direct result of the
ineffectiveness of the water fluidization (only) backwashing tech-
nique which was used to clean the filter.  The test results are also
quite in line with visual observations of the north filter, which
                                  135

-------
OJ
           .O)
o
LU


I
LU

"
o
_i
O
CO
   36

   32


   28

   24


   20
                           O NORTH FILTER - WATER FLUIDIZATION ONLY
                           a SOUTH FILTER - AIR-SCOUR AUXILIARY

                           A WEST FILTER - SURFACE  WASH AUXILIARY
                            10    15     20    25
                                       30    35    40
                                      RUN NUMBER
45
50    55     60    65   70
                      Fig. 33.   Standardized abrasion test results  (Phase II) during direct
                                filtration of secondary effluent.

-------
qualitatively labeled it as the bed in worst physical condition as
described previously.

The lower of the two values plotted for each filter at run 64 was the
abrasion test result at the conclusion of a clean up operation which
will be described in more detail later.  Interestingly, the north
filter was still in worse condition (in terms of solids remaining on
the media) at the end of its extensive clean up procedure than the
other two filters were before the start of their clean up operations.

The abrasion test data is certainly less definitive with respect to
differences in condition of the media between the south and west fil-
ters.  The graphs for these two filters cross each other four times,
although each displays an average upward trend.  The west filter ex-
hibited a sharp peak at run 28, but immediately dropped with the next
test.  This point is viewed with suspicion and is thought to be in
error, although nothing in the test record indicated an error.

Unfortunately, conclusions about differences in the cleanliness of
the media in the south and west filters are not warranted based upon
the available abrasion test data.  It is apparent that the filters
should have been operated for a longer period of time to establish
definite trends; however, this was impossible because of the onset of
winter.

Initial head loss.  Initial head loss readings were recorded at the
start of each run for each filter.  The procedure was standardized by
allowing 15 min after opening the filter influent valve before taking
readings on the piezometer tubes.  The initial head loss, as plotted
for each filter in Fig. 34, represents the total loss in head through
the filter media after 15 min of operation.

By rearranging the data for each filter without regard to time and
plotting the percent of observations exceeding a given head loss
value on normal probability paper, one obtains three curves as shown
in Fig. 35.

Clearly, the north filter exhibits substantially higher initial head
losses than either the south or west filters.  One would expect such
results since throughout the filter run series, the accumulation of
mud balls and agglomerates was most rapid and sustained in the north
filter.  Because of the ineffectiveness of the water fluidization
(only) backwashes, many of these masses were not broken up during the
backwash and remained in the bed at the start of the next run.  Since
much of the bed remained plugged by these solid masses, the filter
influent was forced to find paths around them, causing increased flow
velocity through the cleaner portions of the bed and proprotionately
increased initial head loss through the filter.

Substantial differences in initial head losses between the west and
south filters were not apparent.  Although the west filter data
                                  137

-------
oo
2
§2
                   o NORTH FILTER - WATER FLUIDIZATION ONLY

                   a SOUTH FILTER - AIR-SCOUR AUXILIARY

                   A WEST FILTER - SURFACE  WASH AUXILIARY
                        10

                   Fig. 34.
                              20
40
50
                       30
                  RUN   NUMBER
Initial head loss data for north, south, and west  filters for
entire Phase II study during direct filtration  of  secondary
effluent.
60

-------
         J: 2
          o
         to
         V)


         2

         o
10
o NORTH FILTER - WATER FLUIDIZATKDN ONLY



D SOUTH FILTER-AIR-SCOUR AUXILIARY



£ WEST FILTER - SURFACE WASH AUXILIAR>
                      I
           I
I
I
I   I    I
             0.01    0.1
           1         5    10     20  30  40 50  60  70  80


      PERCENT OF OBSERVATIONS LESS THAN INDICATED VALUE
                                       90    95
                                                99
                       Fig. 35.  Frequency plot of Initial head loss data, Phase II.

-------
 plotted  slightly higher  in Fig.  35  and had nearly  twice  as many ob-
 servations  above 1  ft, the differences were not  appreciable.  For the
 majority of the runs  the initial head loss values  were quite close
 and  followed similar  trends.

 A  typical pattern developed for  all three filters  wherein the initial
 head loss exhibited a cyclic  trend.  Although,there was  no quantita-
 tive means  of correlating initial head loss data with the visual ob-
 servations  of the backwashing procedures, the latter demonstrated a
 similar  trend.  Mud balls and agglomerate accumulations  appeared to
 attain a maximum concentration,  decrease somewhat, and then resume
 their build up.  This trend was  most noticeable  in the north and west
 filters,  but was also observed in the south filter to a  lesser ex-
 tent.  It seems reasonable, although admittedly  speculative, that the
 cyclic trend of initial  head  loss was paralleling  the cyclic solids
 build ups in the filters.

 Head loss development.   Studies  by  Cleasby and Baumann [31] have re-
 vealed that surface layers which form on sand filters used in water
 treatment cause a characteristic exponential shape to the head loss
 versus time curve.  The  accelerated head loss development with time
 was  attributed to the surface layer or "cake" because the layers act
 as filters  themselves.   Once  established, the layers remove  more
 influent  suspended  solids and increasingly compress during the course
 of a run, so that by  the time terminal head loss is attained, the
 layers form dense mats which  are resistant to cleaning.

 As described earlier  in  the section on media appearance, layers of
 organic matter on the filter  surface were frequently observed prior
 to backwashing the  filters in Phase  II.  One might reasonably expect
 an exponential shape  for the  total  head loss curve, although the data
 as plotted  in Figs. 36 through 43 exhibited no such tendency.  The
 nearly straightline development  of  head loss by  the filters could
 have been indicative  of  solids removal occurring much deeper in the
 bed, particularly since  penetration of solids was  consistently visi-
 ble 4 to  6  in. below  the surface in all filters.   However, this does
 not  seem  plausible  since surface layers were observed; therefore, it
 seems as  though some  other phenomenon was acting to alter the shape
 of the curve.

 An explanation of the linear head loss development may be hypothe-
 sized by  observing  the head loss development in  the 4-in. layer of
 media directly below  the surface and by noting that surface cracking
 was observed  primarily in the north and south filters throughout the
 study, as described earlier.  During the course of a run, the bed
 depth was consistently compressed from 1.0 to 1.5  in. in each filter,
which was likely the  result of the  effect of compressible coating on
 the media due to inefficient backwashing.  The compression can cause
 the media surface to crack -and to pull away from the sides of the
 filter housing walls  or  to break, as in a beam failure,  over the pro-
 truding piezometer  tap.  When surface cracks open up as  the filter
                                 140

-------
CO
2
     NORTH FILTER
     (WATER FLUIDIZATION ONLY)
           6.62 ft at 12.75hr
                     0 to
SOUTH FILTER
(AIR-SCOUR AUXILIARY)
                                                               T
WEST FILTER
(SURFACE  WASH AUXILIARY)
                                      3.26 ft at 12.75 hr
                                                     16 to 24 in.
                                                     12 to 24 in.
                                                      4 to 24 in.
                            10    0    2    4    6    8    10
                            TIME FROM BEGINNING OF RUN, hrs
                                               8   10
                Fig. 36.  Chronological head loss development at various media
                         depths, run 2.

-------
10
         I/)
            3
NORTH FILTER
(WATER FLUID IZATION,
 ONLY)

     0 to 24 in. & 4 to 24 in.
                                8 to 24 in.
             0   2
                                12 to 24 in.
                               16 to 24 in.
                                           SOUTH. FILTER
                                           (AIR-SCOUR AUXILIARY)
                                                        0 to 24 in.
                                              WEST FILTER
                                              (SURFACE WASH AUXILIARY)
                                                        0 to 24 in.
 6    8    10    024    68    10   024*
               TIME FROM BEGINNING OF RUN, hrs


Fig.  37.   Chronological head loss development at various media
          depths, run 14.
                                                                             8    10

-------
                NORTH FILTER
      "(WATER FLUIDIZATION ONLY)
               0 to 24 in.
to
CO
O

*
X
  to 24 in.



   to 24 in.


16 to 24 in.
                   SOUTH FILTER
               (AIR-SCOUR AUXILIARY)
                                              0 to 24 in.
                                                        24 in.
                                                        >
                                                        to 24 in.
                                                         12 to 24 in.
                                                       •016 to 24 in.
        WEST FILTER
(SURFACE  WASH AUXILIARY)
                                                                                     0 to 24 in.
                                                         4 to 24in.
                                                                 8 to 24 in.
                                                           «^ol2 to 24 in.
                                                           Q—pQ|6 jo 24 in.
                             10   0    2    4    6    8   10      0

                              TIME FROM BEGINNING OF RUN, hrs
                                                   4
                 8    10
           Fig. 38.  Chronological head loss development at various media depths,  run 22.

-------
             NORTH FILTER
      (WATER FLUIDIZATION ONLY)
      0 to 24 in
jj 5
1
to
CO
o
                4 to 24 in.
                8 to 24 in.

                12 to 24 in.
                                 SOUTH FILTER
                             (AIR-SCOUR AUXILIARY)
                                                  0 to 24 in.
                                    4 to 24 in.
                                                    24 in.
                                                   12 to 24 in.
                                                       24 j.
     WEST FILTER
(SURFACE WASH AUXILIARY)
                                                                               >0 to 24 in.
                                                                 4 to 24 in.
                                                                     8 to 24 in.
                                                                        to 24 in.
                                                                    916 to 24 in.
                        8    10   0    2    4    6    8    10
                             TIME FROM BEGINNING OF RUN. hts
                                                                         8   10
      Fig.
39.  Chronological head loss development  at various media depths,  run 29.

-------
O
oo  o
^)  O
2
Q
             NORTH FILTER
       (WATER FLUIDIZATION ONLY)
               0 to
                         to 24 in.
    SOUTH FILTER
(AIR-SCOUR AUXILIARY)
                                               0 to
                                                         24 in.
                                                         24 in.
                                                        -o] 2 to 24 in
                                                         016 to 24 in
       WEST FILTER
(SURFACE WASH AUXILIARY)
                                            0 to 24 in.
                                                   2 to 24 in.
                                                   6 to 24 In.
                 68    10       02468   10          024
                                  TIME FROM BEGINNING OF RUN, hrs
                Fig. 40.  Chronological head loss development at various media
                          depths, run 36.
                                            8   10

-------
*.  4
 o
             NORTH FILTER
      (WATER FLUIDIZATKDN ONLY)
             0 to 24 in.
             12 to 24 in.
               16 to 24 in.
    SOUTH FILTER
(AIR-SCOUR AUXILIARY)
                                                  0 to 24 in.
                           I
                                                         to 24 in.
                                                            WEST FILTER
                                                    (SURFACE  WASH AUXILIARY)
                                                                              0 to 24 in.
               24 in.
                   to 24 ii
                Pl6 to 24 i
                                                                         in.

                                                                         to 24 in.
                                                                          2 to 24 in.
                                                                       1°16 to 24 in.
                      8   10
                                          8   10
                  02468   10        0246

                  TIME BEGINNING OF RUN, HR

Fig. 41.   Chronological head loss development at various media depths,  run 43.

-------
 _
o
it
t/T
i/-»
2
a
      _        NORTH FILTER
      (WATER FLUIDIZATION
                     " ONLY)
           0 to 24 in, "
                       24m.
          24 in.
      o-o-
         •o-
       12 to 24 in.
•O—O—O—-o
       16 to 24 in.
             I
                             SOUTH FILTER
                          (AIR-SCOUR AUXILIARY)
                              Oto24 in.
in.
 24 in.
 to 24 in.
      WEST FILTER
(SURFACE  WASH AUXILIARY)
             i
             0 to 24 in.
         4 to 24 in.
             8 to 24 in.
             12 to 24 in.
             16 to 24 In.
                        802468    10        024
                                  TIME FROM BEGINNING OF RUN, hrs
                                                                       8
                    10
         Fig.  42.  Chronological head loss  development at various media depths, run 54.

-------
•P-
oo
                       NORTH FILTER
               '(WATER FLUIDIZATION ONLY)
                         0 to 24 in.
         I  5
         V
     SOUTH FILTER
(AIR-SCOUR AUXILIARY)


        0 to 24 in.
                                                            24 in.
                                                            24 in.

                                                             12 to 24 in.
                                                             16 tp 24 ?p.
       WEST FILTER
(SURFACE  WASH AUXILIARY)



            0 to 24 in.
                                             24 in.
                                             >
                                             to 24 in.
                                             12 to 24 in.
                                               to 24 in.
                                  8   10    0246     8100
                                       TIME FROM BEGINNING OF RUN, hrs
                                              8    10
                  Fig. 43.   Chronological head loss development at various media depths, run 64.

-------
run progresses (as they did in this study) they may preclude the for-
mation of an exponential total head loss pattern since the influent
is being introduced deeper into the bed through the cracks.  The net
effect would be to increase the head loss in the media layer directly
below the surface layer, which was frequently observed in the north
and south filters.

In some cases, a crack must have formed by beam failure over the
first protruding piezometer tap at 4-in. depth, which caused the head
loss curve at that depth (4 to 24 in.) to converge on the curve for
the piezometer above the media (0 to 24 in.).  This is evident on
several of the curves for the north filter, which had the most ex-
tensive surface cracking.

Table 14 illustrates the initial total head loss and the time re-
quired to reach 3 ft and 4 ft of total head loss for each filter.  It
is apparent that the initial head loss was higher for the north fil-
ter, which reduced the head available for solids accumulation during
the filter run.  Consequently, the average run length for the north
filter was reduced compared to the other filters.  Differences be-
tween the south and west filters are not great or consistent, so
conclusions are not warranted.

Water quality.  The means and standard deviations of the various
water quality parameters are given in Table 15.  No statistical com-
parison between the filters was attempted for the reasons explained
in Phase I.  There is no apparent difference between the filters,
even though the condition of the north filter was substantially
poorer than that of the other two filters.

As the result of sampling difficulties, composited samples of the in-
fluent were not made for the majority of the observation runs.
Therefore, reporting removal efficiencies on the basis of influent
grab samples was not considered meaningful or proper.  Fortunately,
the Ames Water Pollution Control Plant personnel had conducted sus-
pended solids and BOD determinations on composited samples collected
over a period corresponding closely to the observation runs of this
study; these are included in Table 15.

At roughly 2-hr intervals during the observation runs, grab samples
were collected from the effluent of each of the filters and from the
influent, and turbidity determinations were made.  These values were
treated as a hand-sampled composite, averaged for each run, and re-
ported in Table 15.

Filter clean up operations.  At the conclusion of the regular filter
run series in Phase II, special cleanup operations were conducted on
each of the three filters.  The purpose of these operations was to
determine if the media in each filter could be restored to its origi-
nal state by a series of consecutive backwashes using varied tech-
niques.  Core samples were taken at the beginning and at the end of
                                 149

-------
Table 14.  Summary of head loss development during observation runs  of  Phase II, during direct fil-
           tration of secondary effluent.
Run
2
14
22
29
36
43
57
64
Avg
Initial
N
1.00
2.53
2.58
2.77
1.55
2.19
4.76
1.95
2.42
head lossa
S
0.33
0.62
0.54
1.31
0.81
0.89
0.95
0.84
0.79
, ft
W
0.38
0.96
0.79
1.05
1.09
0.63
0.95
0.88
0.84
Time to 3
N
5.8
1.8
1.5
0.6
4.2
1.6
2.4
2.0
2.5
ft head
S
11.8
5.2
6.3
2.6
6.0
4.8
3.5
3.0
5.4
lossa, hr
W
9.6
5.4
7.2
4.0
5.6
5.8
2.9
3.2
5.5
Time to 4
N
8.4
3.8
3.6
1.6
6.4
3.0
3.6
3.6
4.2
ft head
S
15. 6b
7.4
8.8
4.0
8.3
6.8
5.2
4.0
7.5
lossa, hr
W
12.8
7.8
8.8
6.0
8.2
7.8
3.6
4.6
7.5
 Total head loss across the filter media in each case.

 Extrapolated value.

-------
Table 15.  Results of analyses during direct filtration of secondary effluent (Phase II) from August
           30 to October 30, 1973.  (All composite samples except as noted.)

Suspended solids (mg/1)
cr (N = 8)
Turbidity (FTU)
a (N = 9)
BOD5 (mg/l)b
a (N = 8)
TOG (mg/1)
CT (N = 7)
Total P04 (mg/1)
CT (N = 8)
Ortho PC>4 (mg/1)
cr (N = 8)
Total Kjehldahl nitrogen (mg/1 as N)
(T (N = 8)
Filter
influent
30. 5C
7.6
17.4
2.2
40. 1C
15.4
17. 4d
4.1
20.6d
3.9
19.3d
3.3
12. 7d
4.4

N filter, water
f luidization
only
4.05
2.80
4.84
1.33
12.1
3.1
12.7
2.8
23.9
4.5
22.7
3.7
15.4
2.6
Filter effluent3
S filter,
air -scour
auxiliary
3.63
1.58
4.69
1.42
14.2
4.9
12.0
2.9
24.6
3.3
22.8
2.4
15.5
2.6

W filter,
surface wash
auxiliary
4.10
1.6
4.61
1.36
12.7
4.6
14.2
5.3
24.4
3.7
22.5
2.5
15.1
2.6
 Average filtration rate 1.8 gpm/sq ft

 Nitrification not suppressed in BOD test except in the samples run by Ames WPG Laboratory.

'Two of the observations are from grab samples, and two from composites run by Ames WPC Laboratory.

 Four of the observations are from composite samples, the remainder from grab samples.

-------
 the  clean up procedure  to  evaluate  the overall  effectiveness of  the
 operations.  Additionally,  samples  of the backwash wastewater were
 taken at  30-sec  intervals  and  composited throughout  each  step of the
 three clean up procedures.  The  results of  these procedures are  sum-
 marized in Table 16  and discussed in detail below.

 Initial head loss readings, abrasion test results and visual observa-
 tions all indicated  the media  of the north  filter to be in the worst
 condition of the three  filters.  A  visual check just prior to the
 start of  the clean up operation, further confirmed the poor condition
 of the north filter.  The  media  had a heavy surface  layer, and 1/4 to
 1/2-in. wide cracks  of  varying lengths were visible  at the surface.
 The  sand  and coal layers had become highly  intermixed as  the result
 of poor fluidization and poor  stratification in previous  backwashes.
 Finally,  two  6-in.  diameter agglomerates were  clearly visible.

 To initiate the  clean up procedure  for the  north filter,  the media
 was  subjected to a normal  water  fluidization (only)  backwash.  At
 the  beginning of the fluidization of the media, the  top 6-in. layer
 of anthracite rose as a plug and then settled to the sand-coal inter-
 face.  The backwash was successful  only in  breaking  up this layer
 into large chunks and could not  effect further  breakdown.  A core
 sample was taken at  this stage,  and the abrasion test results indi-
 cated a high value of 33.7 mg/g.  This initial  step  was necessary to
 place the media  in the  state in  which it would  normally be at the
 conclusion of its standard backwashing procedure and, therefore, to
 establish a starting point for the  evaluation of subsequent and
 varied backwashes.

 The  second step  in the  procedure consisted  of a subfluidization  back-
 wash with simultaneous  air scouring, the only introduction of air to
 this media since  the beginning of the Phase II.  Prior to the back-
 wash,  the water  level in the filter was drained to the media surface.
 Air  was then applied at 3.72 scfm/sq ft and water at 11 gpm/sq ft,
 until the water  level reached  an elevation  6 in. below the washwater
 trough.   At this  point  the air was  shut off and the  water rate in-
 creased to 20 gpm/sq ft, which fluidized the media by itself.  The
 combination air  and water wash seemed to produce substantial agita-
 tion in the media and effectively disintegrated many of the mud  balls
 in the media.  However, many mud balls were still present during the
 5-min water (only) backwash which immediately followed, and one
 large, 5-in. mud  ball was observed  deep in  the  sand  layer near the
 conclusion of the wash.

The  third clean up step was essentially a repeat of  the previous
 step, and it continued  to improve the condition of the media.  During
 the combined air  and water wash, the water  rate was  reduced slightly
 to 10 gpm/sq ft,  but the agitation  of the media was  quite good.  The
combined  action was particularly effective because the media exhibit-
ed no tendency to "pack" after 30 sec to 1 min of operation,  as  it
did when using air scour alone as on the south filter during the
                                  152

-------
Table 16.   Data summary  of clean up  operation at  end  of Phase  II.
Steps
Brief description of cleanup procedure - see text
 Solids        Media
released      abrasion
(g/sq ft)       (mg/g)

1.
2.
3.
4.
5.
6.


1.
2.
3.
4.


1.
2.
3.
4.
5.
6.

North Filter
Normal water (only) backwash.
Air and water combination (water at subsidization rate)
as water rose from 1 in. above media to 6 in. below trough.
Follow with water (only) backwash.
Repeat Step 2 with a reduced wash rate during air and
water combination.
Repeat Step 3 except continue air scour after combination
air and water. Follow with water (only) backwash.
Air and water combination, water above fluldlzation.
Follow with water (only) backwash.
Repeat Step 5, but leave air on 3 min. after air water
combination. Follow with water backwash.
Total
South Filter
Normal air (only) followed by water (only) sequence.
Combination air and water with water just at fluidiza-
tlon rate. Follow with air (only) (2 min.) and then
water (only).
Repeat Step 2.
Repeat Step 2.
Total
West Filter
Water (only) backwash.
Subsurface wash and water wash simultaneously.
Repeat Step 1 without drawdown of water above media.
Repeat Step 2.
Air and water combination, follow by 2 min. air (only),
and 'finally 3.5 min. water (only) backwash.
Repeat Step S, except for 5 min. water (only) backwash.
Total

33.7
314
163
205
84
67 15.3
833 - (35.1mg/g)a

9.0
100
42
_17 1.9
161 - (6.8 mg/g)«

11.3
52
18
12
120
_26 2.9
228 - (9.6 mg/g)a
*Based on 12 In. coal media Involved in filtration and weight of coal media of 23,742 g/sq ft.
                                          153

-------
 series.  Layer  stratification was more  pronounced  during  the water
 fluidization backwash which  followed, and  the number  of mud balls  was
 further  reduced.  However, some mud balls  persisted at the sand-coal
 interface,  and  one  large, 5-in. diameter agglomerate  was  observed
 near  the end of the water (only) backwash.

 The fourth  cleaning procedure consisted of  the previously described
 air and  water combination followed by air  (only) for  2 min at  the
 same  rate and concluded with water backwash at 20  gpm/sq  ft.   Careful
 observation during  the combined action  revealed  that  violent agita-
 tion  took place mainly in the upper 18  in.  of the  bed while the  lower
 12 in. became fairly packed  with little movement.  As expected,  the
 agitation during  the air (only) scour was  fairly good for the  first
 minute until the  bed packed, but overall agitation appeared much less
 effective than  the  combined  action.  One large,  6-in. diameter mud
 ball  and several  smaller ones were observed at the conclusion  of the
 water backwash.

 The final two steps in the north filter clean up operation were  es-
 sentially the same  and differed from the fourth  step  only in that  the
 water rate  used during the combined air-water scour was increased  to
 14 gpm/sq ft.   However, this change in  rate caused the media to  be
 fluidized and improved the media agitation  by extending the action
 throughout  the  bed.  By the  end of the  water backwash of  the fifth
 step, the remaining agglomerates had disappeared,  although two
 "clusters"  of four  to six  1-in. diameter mud balls persisted  and  were
 not broken  up even  at the conclusion of the final  step.   In each of
 the last two steps, the air  (only) scour produced  the now typical
 result:  good action during  the first minute but little movement
 after 1  min because of intermixing and  packing of  sand and coal
 layers.

 Throughout  the  regular filter run series, the south filter was
 cleaned  using a 5-min air-scour auxiliary at 3.72  scfm/sq ft followed
 by a  5-min  water  fluidization backwash  at 20 gpm/sq ft.   This  proce-
 dure was used to  initiate the cleanup operations on the south  filter.
 The media was not in nearly  as poor condition as the  north filter,
 but during  the  last 30 sec of the water wash, six  3 to 4-in. diameter
 mud balls were  observed.  At the conclusion of the initial step, a
 core  sample was taken and subjected to  an abrasion test.  The  test
 results  indicated the media  to be approximately  four  times cleaner
 then  the north  filter at the same stage.  The clean up procedure
 consisted of an air and water combination wash with the media  slight-
 ly fluidized (14 gpm/sq ft), an air (only)  scour for  2 min at  3.72
 scfm/sq  ft,  and a water (only) backwash at  18 gpm/sq  ft.  Expanded
 bed depth remained constant  at 32-1/2 in. for the  remaining three
 steps, and fluidization and  stratification were  consistently good.
As the clean up continued,  the number of mud balls observed during
 the water backwashes diminished.  At the conclusion of the fourth  and
 final step,  only one small mud ball was observed,  and the physical
 appearance of the bed was excellent.   A final core sample taken  at
                                 154

-------
this time was analyzed in the laboratory and was found to verify the
visual observations:  the abrasion test value was determined as 1.88
mg/g which indicated the bed was in very good condition.

As one will recall from an earlier discussion, the west filter had
been equipped with a rotating surface washer which operated both in-
dependent of and in conjunction with the water backwash during the
backwash sequence of the filter series.  The routine wash was not
used to begin the west filter clean up operation; instead, a water
(only) backwash rate of 19.5 gpm/sq ft for 5 min was used.  The sur-
face washer had been modified prior to the start of the wash, and
while the media was fluidized, the rotating washer was pushed to the
fluidized interface of the sand and coal layers for use in subsequent
steps.  At the conclusion of the water wash, a core sample was taken,
and a subsequent abrasion test resulted in a value of 11.26 mg/g,
very close to that of the south filter.

Steps 2, 3, and 4 were combination rotating subsurface wash, water
backwash, with the former operating for 3, 4, and 3 minutes, respec-
tively.  The rotating washer had been lowered into the bed to see if
it could effectively break up the mud balls which accumulated at the
sand-coal interface.  By the end of the second step no mud balls were
seen even though several large (4 to 8-in. diameter) mud balls had
been observed at the start of step 2.  Only three small mud balls,
approximately 1 in. in diameter, were seen during the third step, and
none at all were observed during the fourth step — the last combina-
tion subsurface wash, water backwash used in the cleanup.  The rotat-
ing subsurface washer, therefore, appeared to do an excellent job of
breaking up the large agglomerates and mud balls in the media.

At the conclusion of the fourth step, however, there were signs that
the media was not thoroughly clean even though the mud balls had been
eliminated.  Dirt remained floating on the surface of the coal after
step 4, and the anthracite itself had a grayish cast, as though it
were still coated.  This seemed an excellent opportunity to change
the backwash procedure for step 5 to include the use of air scour and
see if it could further clean the coal layer.

The procedure for step 5 was, therefore, changed to provide combina-
tion air and water wash with the water wash rate set at 12.5 gpm/sq
ft to fluidize the media as the water rose from the media to near the
overflow.  This was to be followed by 2 min of air (only) scour at
3.72 scfm/sq ft and, lastly, by 5 min of water (only) backwash at
20 gpm/sq ft.  Prior to step 5 the composited backwash waste water
had shown a decline in suspended solids concentration equivalent to a
dirt released value of 12 g/sq ft after step 4.  The composited sam-
ple for step 5, however, indicated a nearly ten-fold increase in the
dirt released value to 120 g/sq ft.  These results seemed, at first
glance, even more significant because a dwindling supply of backwash
water limited the actual water (only) backwash to 3-1/2 min.
                                 155

-------
 Step  6  in the west filter clean up was identical to step 5, although
 the water (only) wash was extended 5 min  after building up the back-
 wash  supply.  It was noted that,  again, the composited backwash waste
 sample  had  a suspended solids concentration twice  that of step 4.
 During  this final step of the west filter cleanup, no mud balls were
 observed, all floating material had disappeared from the surface, and
 the anthracite regained its rich, black lustre.  An abrasion test of
 a core  sample taken after this step was completed  yielded a value of
 2.19  mg/g, which further confirmed the clean condition of the bed.

 The results from steps 5 and 6 initially  appeared  to demonstrate the
 superiority of the air and water  combination wash  over the rotary
 surface wash backwash auxiliary.  However, as noted in step 4, much
 dirt  was visibly released from the media  which was not carried out
 during  the water (only) backwash, and large solids were also noted
 adhering to the sides of the filter housing.  At the conclusion of
 step  6  the floating material had  disappeared entirely, and some of
 the wall solids had also been removed.  Therefore, one can state con-
 fidently that the air and water combination backwashes (steps 5 and 6)
 were  definitely more effective in loosening solids from the filter
 housing, and in disintegrating solids so  they could be transported
 out of  the filter during the water (only) backwash, but one cannot
 state categorically that it was more efficient in  releasing dirt from
 the media.

 The data presented in Table 16 clearly demonstrate  the benefit of
 both  auxiliaries in maintaining the filter media in cleaner condi-
 tion.   However, it is not possible to choose which auxiliary is bet-
 ter from the data.  The fact that the air and water used simultane-
 ously were able to release substantial additional  solids from the
 south and west filters, which had been routinely washed with air-
 scour and surface wash auxiliaries, respectively,  implies the superi-
 ority of air and water together as a backwash method.  One should be
 careful about jumping to such a conclusion, however, since the re-
 search  of Phase II was not designed to prove that point.  To prove
 the superiority of air and water  together over the other two backwash
 auxiliaries, it would be necessary to conduct an entire research
 phase in which the three methods  of backwash auxiliary were compared
 in parallel.

 Summary and Conclusions - Phase II

The objectives of the experimental investigations in Phase II were to
determine the effectiveness of three different backwashing techniques
on dual-media filters and to compare the performance of the three fil-
 ters while filtering secondary effluent.  The backwashing techniques
used were as follows:

1.   North filter, backwash by water fluidization  alone.
                                 156

-------
2.   South filter, two-phase sequence consisting of air (only) scour
     followed by water fluidization backwash (i.e., air-scour auxil-
     iary) .

3.   West filter, three-phase sequence consisting of surface wash
     (only), surface wash and water backwash, and water fluidization
     backwash.

The experimental data were collected over a nine-week interval at the
Ames, Iowa, trickling filter plant using pilot-scale equipment.  The
following are the conclusions of the phase:

1.   None of the three methods of backwashing was able to keep the
     filter bed completely free of mud balls and agglomerates.

2.   Based upon higher initial head losses, steadily increasing solid
     accumulation observed in abrasion test results, visual observa-
     tions of the media condition, and results of the clean up opera-
     tion, the north filter was clearly in the worst condition of the
     three beds.  Therefore, water fluidization (only) backwashing is
     ineffective in maintaining the filter media in good condition,
     and some means of auxiliary cleaning is required.

3.   On the basis of the items mentioned in conclusion 2, no conclu-
     sive differences were observed between the south and west filter
     in the effectiveness of their cleaning techniques, air-scour
     auxiliary and surface wash auxiliary, respectively.

4.   No apparent differences were observed in the effluent qualities
     among the three filters, particularly in the primary removal
     efficiency parameters BOD, suspended solids, and turbidity.
     This indicates bed condition played little part in removal ef-
     ficiency in this study; however, this may have been due to the
     choice of a fine filter media, low filtration rate or low termi-
     nal head loss in this research.

5.   The use of some form of air-scour auxiliary or some form of sur-
     face wash auxiliary is essential to the satisfactory functioning
     of wastewater filters.  The methods used in this research did
     not completely eliminate all dirty filter problems, but both
     auxiliaries reduced the problems to acceptable levels so that
     filter function did not seem to be impaired.

             Operation and Results - phases III, IV and V

              Single-, Dual-,and Triple-Media Filtration

                         of  Secondary Effluent

Phases III through V were prompted by the deficiencies in water flu-
idization backwashing demonstrated in Phase I and II, even when as-
sisted by air-scour or surface wash auxiliary.  These phases were
                                 157

-------
 also prompted  by  the  implied  superiority  of  simultaneous  air  and
 water backwash revealed  in  the  cleanup  operations  at  the  end  of
 Phase I  and  II.   There were other  questions  raised by Phase II which
 needed evaluation so  Phases III through V were designed to try to
 answer those questions.

 During Phases  III through V,  the north  filter was equipped with dual
 media.   The  backwash  included three  steps, including  air  scour alone,
 air  and  water  simultaneously  for a very brief period  without  over-
 flow,  and  finally, water fluidization backwash alone.  Coarser under-
 drain strainer openings  were  used  which required the  use  of gravel
 below the  media.  A double  reverse graded gravel was  used to  resist
 movement by  the simultaneous  air and water backwash action.

 The  south  filter  was  equipped with a deeper  bed of coarse sand (6 to
 10 mesh, 2 to  3.36-mm sieve range) and  was washed at  subfluidization
 rates with air and water simultaneously during overflow for a rather
 extended period,  followed by  a  brief period  of subfluidization water
 backwash.

 The  west filter was equipped  with  triple  media (i.e.,  commercially
 obtained "mixed-media" from the Neptune Microfloc Corporation).  The
 media was  underlain by a graded gravel, and  the water fluidization
 backwash was assisted by a  surface and  subsurface washer.  The sub-
 surface  washer was added in this phase  because of the problem of mud
 balls sinking  to  the  coal-sand  interface  and floating there out of
 reach of the normal surface washer.

 The  details  of media  sizes  and  depths,  gravel gradation,  and  .ider-
 drain strainers have  been presented previously.  Details  of the back-
 wash are presented in the following pages.

 Operation  -  Phases III through  V

 The  first  filtration  run of the testing period was on May 16, 1974,
 and  operation  continued  daily,  except when the equipment  malfunc-
 tioned,  to the cleanup operations  on November 2, 1974.  The five-
 month period was  divided into three phases coinciding with changes in
 media  and  sampling technique.   The first  runs were designated as
 Phase  III, a continuation of  the notation started the  previous year,
 and  were continued to July 18,  1974, when Phase IV began.  Due to an
 incident of  clogging  of  the uriderdrain  strainers (to  be described in
 detail later)  and the rebuilding of two of the filters, Phase IV was
 started here to signify  these changes.  It continued  to August 27,
 1974, or to  the start of Phase  V.  New  sampling containers were pur-
 chased and a more careful sample bottle washing procedure was used
 starting on  this date, so a phase  change was felt to be appropriate.
Each particular day of filtration was designated as a  run.  Phase III
 ran  from run 1 to run 50; a new run sequence was started  at Phase IV
but not  at Phase V, so from July 18 to  November 2, 1974,  a single
 sequence of  runs was  used (run  1 to run 99).  Rather  than report the
                                 158

-------
results by phase and run number, which is misleading for Phase V
since a new run number sequence was not started, the actual number of
the day of the year the run was performed will be used to designate
the occurrence.  For example, Phase III began on May 16 or calendar
day no. 136 and the cleanup operations were November 2 or day no.
306.  This notation will be used to discuss the results of Phases III
through V.  Another reason for this choice of notation was the method
used to figure averages and to plot data points over the entire test-
ing period.  It was simpler to have the abscissa as one continuing
sequence rather than as various run and phase numbers.

The flow rate for Phase III was 2.1 gpm/sq ft.  The flow rate varied
slightly between runs in Phase III but was evenly split between fil-
ters in any particular run.  At the start of Phase IV and continuing
through Phase V, the flow rate was changed to 3.2 gpm/sq ft (7.8 m/hr),
a 50% increase.  Here again the flow rate between runs was slightly
variable.  The flow rate for the runs discussed will be presented
individually in a later section.

Backwashing.  One objective of this study was to compare the effec-
tiveness of three different backwashing procedures.  One backwashing
procedure was assigned to each filter for the length of the study,
and it was not changed although minor changes in duration of the vari-
out backwash operations were adjusted as they seemed necessary.  Al-
though each filter was backwashed by a different technique, they all
were prepared for backwashing by the same series of steps as was used
in prior phases.

The dual-media filter was designated for an air-scour cycle and a
fluidized waterwash.  After some initial experimenting in the first
few runs, the following backwash procedure was decided upon for the
dual-media filter.

Step 1.  Drain water from above the media to within 1 in. of the sur-
         face.

     2.  Add air at 3 scfm/sq ft (standard cubic ft/min of air at
         70 °F and 1 atmosphere  pressure)  and water at 13 gpm/sq ft
         in combination until 6 in. below overflow trough.

     3.  Shut air off and continue water alone wash at 23 gpm/sq ft
         for 5 min.

There were a few days where the air wash rate was 4 scfm/sq ft, but
the above procedure was run as indicated for approximately 2-1/2
weeks.  At that point, the water alone wash rate was cut to 21.5 gpm/
sq ft, and that change was continued to run 38 (day 183) of Phase III.
The following backwashing procedure was the one used from run 38 to
the end of the Phase V.
                                  159

-------
 Step  1.  Drain down  as before.

      2.  Apply air scour  alone  at  3  scfm/sq  ft  for 2 min.

      3.  Add water at 13  gpm/sq ft in  addition  to the
         air until the water is 6  in.  below  the overflow.

      4.  Shut off air and continue water at  21.5 gpm/sq ft for 5 min.

 The only minor change to  the above procedure came at run 32, Phase IV
 (day  no. 232), when  the time of the  air scour was increased to 5 min
 and continued to the end  of the study.

 The mixed-media filter was equipped  with a surface and subsurface
 auxiliary washer.  The backwash technique for this filter began by
 starting the surface washer and continuing it for 2 min.  The surface
 washer was left on,  and then the backwash water was started at 15
 gpm/sq ft, fluidizing the bed,  and followed  by  the starting of the
 subsurface washer.   The subsurface washer failed to spin at all until
 the bed became fluidized.  After 3 min, the  surface and subsurface
 washers were stopped, thus providing 5 min and  3 min total operating
 time, respectively.  The  water  alone backwash continued for 4 min to
 end the cycle.  Throughout the  first 16 runs of Phase III, the water
 only  backwash lasted for  only 2 min, and the water rate used during
 surface and subsurface washer cycle  varied from 13 to 15 gpm/sq ft.
 However, by run 17,  the backwash cycle first mentioned was established
 and used until run 30 of  Phase  III.  There had been some trouble get-
 ting  the surface washer to break up  the surface mat.  The washer was
 just  a little too high to fully agitate the  surface of the media.
 Beginning with run 30 (day 173), a very small amount of backwash
 water was added during the first 2 min to provide a slight expansion
 of the bed, enabling the  surface washer to do a more effective job.

 The coarse sand filter was backwashed with air  scour and water simul-
 taneously at washwater rates far below fluidization for the media.
 The backwashing sequence  involved  only two steps.  The first step ap-
 plied an air-water combination  wash  at rates of 7 scfm/sq ft and 9.7
 gpm/sq ft, respectively,  for 15 min.  The second step was a waterwash
 only  cycle at the same rate for 5  min  to expel  as much air as possi-
 ble from the media.

 Sampling procedures  and data collection.  Sampling procedures and data
 collection were similar to those described for Phases I and II, with
 the following exceptions.  An influent composite sampler was in oper-
 ation during Phases  III through V  in addition to the effluent samplers
which had been in service during prior phases.

Observation  runs were conducted two times per week during Phases III
 through V.   As before, an observation run consisted of carefully mon-
 itoring and recording of  the performance parameters throughout the
duration of the run.   In most cases, all three filters were observed
                                  160

-------
up to the point when the head loss reached the splitter box elevation
and the filter began bypassing.  The run lengths for the dual- and
mixed-media filters were much shorter than for the coarse sand media.
Therefore, during observation runs, filters were backwashed whenever
they reached terminal head loss and put back in service so they would
operate at constant rate for the same total amount of operating time
as the coarse sand filter.  Detailed observations were not continued
for these two filters after the first backwashing, but the composite
samplers were continued in operation (except during backwashing) to
obtain a sample from each filter over the same time period.

Results - Phases III through V

Visual observations - dual-media filter.  The dual-media filter
(north filter) showed good backwashing in the first stages of the
study.  The filter had been in operation five days, day no. 144, be-
fore the first mud ball appeared, 6 by 1 in.  The small mud balls
grew over the next week until the air-water combination wash failed
to disperse the surface mat.  Since the air-water wash could be run
just until the backwash water was near the overflow trough, the bed
received approximately 1 min of the combination wash.  When the sur-
face mat was thick and highly compacted, this was not enough time to
completely break up the mat.  However, the air-water wash did provide
excellent agitation throughout the entire bed when being applied.
The coal and sand were completely mixed after only the short time the
combination wash was applied.  Complete fluidization was then required
to restratify the filter media.  With the clean filter media in the
first part of this study, the fluidization and stratification was
easily accomplished.  However, as the size and population of the mud
balls increased, jetting action occurred along with dead areas in the
bed.  Sometimes sand remained mixed with the coal in various areas
above excessive mud ball deposits.  A typical comment recorded in the
data book over the next month of operation was 5 to 6 mud balls, 2 to
3 in. in diameter, excellent air-water wash agitation followed by
complete fluidization and restratification.  After over a month of
operation, two successive days of observations indicated no mud balls
present in the dual-media filter.  However, the backwash for day no.
177 failed to break up an 8 by 3-in. agglomerate.  The agglomeration
sunk to the interface, channeling the backwash water around it and
producing a "dead" space above it.  The following backwashes had
similar results with the formation of an increasing number of mud
balls and less effective backwashing.  Water channeling became com-
mon, and cracking was evident 1-1/2 in. into the bed.  During the
next few days, the top 2 to 3-1/2 in. of media became heavily packed
with solids, which made the air-water combination almost totally in-
effective.  The bed was in extremely poor condition on day no. 181,
with numerous dead areas and layers of previous surface mats resting
in the coal.  On day no. 183 the backwashing procedure was modified
in an attempt to improve the condition of the bed.
                                  161

-------
 The  new procedure  added  a  step which  applied  air  only.  The  complete
 backwashing  technique was  now 2 min of  air scour,  air-water  wash  to
 just under overflow, and,  finally, 5  min  of waterwash only.  This
 modification immediately improved  the condition of the bed.  Since
 the  coal was not so agglomerated now, the air-water wash efficiency
 was  improved as well.  It  was noted that  the  action of the air  scour
 alone was most effective during the first 30  sec  of application;
 after that the scouring  action was confined to only the top  2 to  3
 in.  of  media.  The condition  of the bed improved,  as indicated  by the
 decreasing number  and size of mud balls and agglomerates until  on
 day  no.  188,  the comment recorded was,  "...a  few mud balls present
 but  not to any great extent." It would seem  the  addition of the  air-
 scour cycle  improved the bed  condition  of the dual-media filter.  It
 was  at  this  time,  trouble  developed with  the  strainers, and  the fil-
 ter  was rebuilt, signaling the commencement of Phase IV.

 After the filter was rebuilt, the same  backwashing procedure was
 still used.  The first week showed no problems or  mud ball formation,
 rather  complete mixing during the combination wash, good fluidization,
 and  very distinct  restratification.   Small  mud balls (2 to 3 in.  in
 diameter) were observed  the following week  until during the  third
 week of operation  large  pieces of the surface mat  (2 by 6 in.)  were
 observed falling into the  bed.  Day no. 221 and 222 note large  mud
 balls (3 by  4 in. and 4  by 5  in.) falling to  the  interface.  Typi-
 cally,  along with these  large agglomerations, several smaller (1/2 to
 1 in.)  were  noted as well.

 On day  no. 232, a special  backwashing sequence was performed to try
 and  rid  the  dual-media filter of its high surface  layer head loss.
 The  first step of the special backwash  was  simply  the routine back-
 wash procedure combined with  turbidity  measurements of the dirty
 washwater collected at 30-sec intervals.  After completion of step 1,
 it was noted there was a considerable amount of gray organic matter
 on the  surface of the filter  along with some 2 by  3-in. mud  balls 8
 in.  below the surface.   Step  2 of the special backwash was:  (a) 5
 min  of  air alone, (b) air-water wash for  approximately 30 sec as the
 water rose to the overflow trough, and  (c)  5 min of water alone, all
 at the  same  rates as previously used.   Six  small (1 by 2 in.) mud
 balls were observed following Step 2.   Following this day, the  rou-
 tine  backwashing sequence was changed to  that of Step 2.

 On day no. 235 it was noted that the top  8  in. of  fluidized coal
were  individual grains covered with a hairy,  stringy slime.  When
 f luidization was stopped,  this 8 in.  of media was  compacted  into  the
 top 2 to 3 in. of media,  causing excessive  initial head loss.   Spe-
cial observations were made on day no.  241 which described the  string-
ers made of organic matter, 1/16 to 1/8-in. long,  and fuzz-like in
appearance.   A special backwash was conducted on day no.  242 to at-
tempt to correct this problem.
                                  162

-------
The special backwash consisted mainly of simultaneous air-water wash
at low rates for an extended period of time.  The media was subjected
to 1.5 scfm/sq ft of air and 7.5 gpm/sq ft of water for 21 min.  The
media that was carried out of the filter was collected in a garbage
pail to be replaced later.  Samples of the waste backwash water were
collected every minute.  Figure 44 presents the suspended solids and
turbidity for the samples versus the time they were taken.  The
stringy growths were reduced but not eliminated.  The initial head
loss for the two days prior to this experiment was 1.19 ft and 1.28
ft, while after this day it was 0.86 ft.

The hairy growths were still a problem five days later, so another
special backwash was conducted.  The lost media was again collected
and returned, and the bed was then restratified.  No samples were
taken, but these.observations were made.  The stringly attachments
did not appear as long or thick as before; however, the initial head
loss following the experiment was 1.38 ft.  No positive changes were
seen in the condition of the dual-media filter, so chlorination was
tried.  The filter was drained down to within 1 in. of the surface
and 200 ml of household bleach (5.257. sodium hypochlorite by weight)
was added, stirred in by air scour, and let set for 10 hr before
backwashing.  No immediate results were seen.  It was at about this
time, day no. 255, that a similar stringy coating was noted on the
mixed-media filter but in a much milder concentration.  This condi-
tion was recorded for only a few days for that filter, and day no.
269 was the last day the stringy attachments were mentioned on
either of the two filters.

Following the disappearance of the stringy attachments, the operation
of the dual-media filter was rather routine to the end of the testing
period.  The air scour alone provided good mixing and bed agitation
for the first minute, then the air passed upward through the bed in a
channelized manner, causing little action, except as the air bubbles
came out of the bed surface, which violently mixed the top inch of
media.  The air-water wash accomplished excellent mixing, but the
average duration was only 30 sec.  Typical mud ball observations were
4 to 5 in number and 1 to 3 in. in size.  Every few days the surface
mat would be extremely thick, causing pieces of it to tumble down to
the sand-coal interface where they would disintegrate or reduce in
size.

Visual observations - mixed-media filter.  As previously stated, the
mixed-media filter was equipped with a surface and subsurface washer.
The main observation which relates to mud ball formation and bed con-
dition was to note if the washers were working properly.  Following
one day of operation, the surface washer turned freely with a 30 psig
line pressure.  The subsurface washer did not turn, but the jet ac-
tion of the washer nozzles lifted coal layers within the bed.  The
next week of operation caused excessive mud ball formation, still
without the turning of the subsurface washer.  On day no. 146, a
booster pump was installed on the water supply to the surface and
                                  163

-------
300-
RESULTS OP SPECIAL BACKWASH
DAY NO. 242
o SUSPENDED SOLIDS
   TURBIDITY
               80

               70


               60


               50


               40


               30


               20
                                                          A   /BACKWASH"]10
                                                          0   I  SUPPLY
                                     I
  I
I
0
                    6    8     10   12    14    16
                     DURATION OF BACKWASH, min
            18
20   22
           Fig.  44.  Results of special backwash, day no. 242.

-------
subsurface washer line to increase the line pressure and possibly
free the "hung-up" washers; the line pressure was increased to 60 to
80 psig.  This did not solve the problem, and the following day a
large 3 by 4-in. agglomerate and 10 to 12 small,1-in. mud balls were
observed.  These mud balls fell to the interface and were still not
broken up because the subsurface washer was below the interface.
Finally, the rubber caps on the subsurface washer were removed and
the washer rotated by hand, freeing the washer so that it worked
properly for the next two weeks of operation.

The action of the surface washer was described in detail for several
observations.  The washer violently mixed the top 3 to 4 in. of the
coal.  Ineffective action was noted in the corners of the filter
housing, as could be expected with a rotary washer in a square
housing.

Sometimes the surface cake was broken up by this mixing action, while
at other times it simply broke into smaller pieces which, upon fluid-
ization, fell into the bed.  Also during this same period, considera-
ble amounts of fine silica and garnet sand were observed working
their way into the coal layer.  The condition of the bed gradually
deteriorated until channeling of the backwash water was common and
numerous mud balls, 4 to 6 of 2 to 3 in. in diameter, were observed.

On day no. 173, a backwashing procedure change was made to increase
the effectiveness of the surface washer.  During the surface washer
only cycle, a small amount of backwash water was now added to expand
the bed so the surface washer was submerged in the top 1 to 2 in. of
media.  At the low rate of application, the backwash had trouble
raising the thick, heavy surface mat to the surface washer, so the
ineffective wash was still present at times.  The surface washer
worked best when the bed was completely fluidized, but since several
large mud balls (5 by 2 in., 4 by 2 in., 10 by 2 in.) were observed
on day no. 180, complete cleaning of the bed was still not accom-
plished.  Shallow cracks were observed at the top of filter media.
Solids penetration into the bed in general was 2 to 4 in.  The lower
12 in. of the coal contained silica and garnet sand and was highly
agglomerated on day no. 191, just before the filter was rebuilt.  Mud
balls and agglomerates were so dense that complete fluidization and
stratification was not attainable.

The media depths were changed during the rebuilding so the washers
were relocated to better coincide with the interface and surface.
Also, modifications of the washers themselves were made, with changes
in the washer arms and nozzles.

The line pressure was commonly 60 to 80 psig after the modifications,
which helped break up mud balls and surface mats in the first few
runs.  Typically, both the surface and subsurface washers worked well
at times if all the water was channeled through each one separately.
However, both washers would not turn at the same time unless almost
                                  165

-------
 all of  the water went  through the  subsurface washer; this condition
 required  a very delicate balance and was not reproducible from day
 to day.   The bed remained in good  condition for the first two weeks;
 no mud  balls were observed on day  no. 210 and 211.  Both the washers
 worked  fine in the following days, and on. day no. 216, pieces of the
 surface mat that were  not destroyed by the surface washer, fell and
 were broken up by the  subsurface washer.  One or two mud balls of 1
 to 2 in.  in size were  typically observed during this period of opera-
 tion.

 For the next several weeks of operation, the equipment worked fine,
 and a very few small mud balls were observed.  Smaller mud balls (less
 than 1  in. in diameter) were seen  floating between the washers, too
 small to  fall down and be broken by the subsurface washer.  No cracks
 were observed, and penetration of  solids was estimated at between 4
 and 5 in.  The media was reported  to be very clean for day no. 235,
 244, and  247.  Both the fluidization and stratification were easily
 accomplished with the  clean, almost mud ball-free media.  It was at
 this time, day no. 256, that the stringy attachments similar to those
 observed  on the dual-media filter were noticed.

 A special observation was performed on day no. 241 to document in
 detail  the operational characteristics of the washers.  When the
 surface washer was on by itself, the line pressure was about 70 psig
 with the  washer rotating at approximately 80 rpm.  A steady-state
 revolution condition was hard to maintain when both washers were
 operating; 27 rpm was  typical for both.  The subsurface washer alone
 operated  at 75 psig and 27 rpm.  The surface washer provided violent
 agitation to the top 6 in. of media when it was immersed in the flu-
 idized  bed, but the water fluidization (only) cycle provided very
 little  action.  On day no. 263 trouble with the subsurface washer was
 noted;  only 20 rpm was observed.  The surface washer revolved at 100
 rpm but was throttled down to 60 rpm for the backwash.

 The first night in October, day no. 274, was extremely cold, causing
 the surface washer booster pump to freeze and split.  A used pump was
 installed, but the washers turned  at a slower rate thereafter, and it
 was feared dirt particles in the replacement pump plugged a portion
 of the washer nozzles.  Because the surface washer turned much slower
 now and sometimes not at all, mud balls began to develop in the bed.
Four to five, 2-in. and 10 to 12 less than 1-in. mud balls were com-
mon.  In  the remaining month of operation several days of no washers
 rotating or only one of the two washers rotating were noted.  The
bed condition quickly showed signs of the loss of the auxiliary
washers.  Finally, on day no. 293, a length of pipe was used to manu-
 ally reach into the filter and turn the washers.  The surface washer
was freed quickly, but the subsurface washer failed to turn at all.
 In the process of trying to free the subsurface washer, the bed was
 fluidized and stirred for approximately 30 min, breaking up all the
                                166

-------
mud balls present.  The subsurface washer failed to turn again for
the remainder of the study.

Visual observations - coarse sand filter.  As previously stated, the
coarse sand filter was backwashed for a longer period at wash rates
below minimum fluidization velocity.  The backwashes were all very
similar and rather "dull" as compared to the other two filters.  No
mud balls were formed, fluidization did not occur, and extensive
descriptions of the backwash sequence were not warranted.  Very sim-
ple, short data notes completely described all the action taking
place.  The first backwashes provided only 5 min of air-water wash
followed by 3 min of water only.  The dirt could be seen moving up-
ward and out of the bed.  The typical backwash pulsed the top 6 to
8 in. of the bed.  Pronounced air channeling causing jetting which
mounded the top couple of inches of media and produced an uneven
surface for the rest of the study was noted.  The mounds moved about
during successive backwashes.  Early in the study, the air-water wash
was changed to 15 min in duration and the waterwash to 5 min.  Nine
days into Phase III, the bed began to show signs of dirt accumula-
tion.

Dark grey areas were starting throughout the bed except in the top 4
in., which retained their original appearance.  The pulsing action was
noted in the top 12 to 18 in. as streaks of dirt were seen on the
plexiglass window being carried away.  The bed cleaned up nicely
after each wash, particularly the top 18 in.  On day no. 157, darker
areas appearing in the bed were reported.  A 2 to 3-in. vertical
strip along both edges of the window in the corners of the filter
box was covered with this matter.  The dark patches at the window
also started to develop at the two-thirds depth and below.  On day
no. 165, the dark areas shifted to the lower parts of the filter,
covering the bottom 15 in. of media.  Penetration of solids into the
filter was recorded between 12 and 15 in. on day no. 185.  The filter
condition remained very constant for the remaining four months of the
testing period.  The bed cleaned very well with the pulsating action
except for the bottom 15 in. of the filter, scattered areas of anaero-
bic deposits which shifted position as the filter runs progressed,
and areas along the corners of the filter box.

A special investigation was conducted to determine if the bottom 15
in. of media showed any progressive increase in initial head loss
due to its dirty condition.  Figure 45 shows the head loss in the
bottom 16 in. of the coarse sand filter.  As can readily be seen, no
distinct pattern was indicated.  The head loss increased when the 50%
rate increase was made at the start of Series IV.  No definite con-
clusions can be drawn about this dark 15 in. in the bottom of the
filter.  A possible explanation could be simply a wall effect in con-
junction with the strainers which were located at about 4 in. from
the window and 10.5 in. apart.  They created a dead space near the
walls and between the strainers which received ineffective washing.
                                 167

-------
00
 o
£
 «i
O
o
0.40
0.35
0.30
0.25
0.20
0.15
0.10
0.05
 0.0
                    -  o
                                                               INITIAL HEAD LOSS IN BOTTOM
                                                               16 in. OF COARSE MEDIA FILTER
                                 PHASE III
                                2.1 gpm/sq ft
                                 «— PHASE IV  >k  PHASE V
                                 <— 3.2 gpm/sq ft
                                                             O   O   O
                                                             ~      O
                                                       8
                                                                 o
                                                                o   o
                                                                        °0
                                    
-------
The visual observation and operation of the coarse sand filter was
very routine.  However, very late in the study, the left and lower
half of the bed became dirtier.  By the growth of the dark areas it
was obvious the left strainer near the window was partially or fully
plugged.  The study was terminated due to cold weather without any
further changes.

Abrasion tests.  The abrasion test was also used in Phases III
through V as a direct measure of backwashing effectiveness.  The re-
sults of the abrasion tests for the entire test period cannot be com-
pared because two different tests were performed.  As previously
described, Phase III employed a complex, rather long abrasion test
procedure while Phases IV and V used a short modified version of the
same basic test.  For that reason, the abrasion test results for
Phase III are presented in Fig. 46 and the results for Phases IV and
V in Fig. 47.

An abrasion test was performed on the clean filter media to obtain a
control or background level to see if the media itself was being
abraded.  An unused sample of both the anthracite coal and coarse
sand was subjected to the longer abrasion test used for Phase III.
First, however, the coal was mixed for 45 min with the abrasion pro-
peller and then flushed with water until the rinse water remained
clear.  The sand sample was subjected to only the rinsing water be-
fore testing.  The mixing and rinsing of the filter media prior to
the abrasion test was to remove any dust remaining on the media from
shipment.  The coal and sand samples were then subjected to the abra-
sion test procedure.  The results for the new media using the Phase
III procedure were 1.38 mg/g for coal and 0.94 mg/g for sand.  Sever-
al days later, the same media sampler were again subjected to the
abrasion test procedure, yielding 3.17 mg/g and 0.39 mg/g for the
coal and sand, respectively.  It was this high value, 3.17 mg/g, for
the coal that prompted the change of the abrasion test procedure in
Phase IV.  Too much abrasion of the coal itself was taking place
during the 30-min mixing period of the Phase III test procedure.
Since the abrasion testing procedure was modified at the start of
Phase IV, a new standardization was necessary.  Here again, an unused
(out-of-bag) sample of both the coal and sand was mixed and flushed
as before.  The abrasion test results using the new procedure were
0.04 mg/g for coal and 0.018 mg/g for sand.

Despite the fact that the testing procedure was modified during the
testing period, the same pattern can be seen for all three series.
The abrasion test results for Phase III, Fig. 46, shows the coarse
sand filter always had cleaner media than the other two filters.
Also, except for the very first and last values, the mixed-media fil-
ter was cleaner than the dual-media filter.  Furthermore, the graph
shows an erratic buildup of solids for the dual-media filter.  The
coarse sand and mixed-media filters show a rather level result, ex-
cept that there is some erratic behavior in the mixed-media filter in
the latter days of the series.  Based on the abrasion test results of
                                 169

-------
   16.00



   14.00



   12.00
•s
 s 10.00
o
LLJ
2  8.00



    6.00
CK

CO
 O
    4.00



    2.00



    0.00
                      ABRASION TEST RESULTS

                             PHASE III

            O  DUAL MEDIA - AIR-SCOUR WASH

            A  MIXED-MEDIA - SURFACE WASH

            D  COARSE MEDIA - SUBFLUIDIZATION WASH,
        150
                    160           170

                         CALENDAR DAY, 1974
180
190
        Fig.  46.  Standard abrasion test results for Phase III.

-------
          ABRASION TEST RESULTS
              PHASES IV & V
O  DUAL MEDIA - AIR-SCOUR WASH
   MIXED-MEDIA-SURFACE WASH
   COARSE MEDIA - SUBFLUIDIZATION WASH
                      245     255      265
                       CALENDAR DAY,  1974
 Fig. 47.  Standard abrasion test results  for Phases  IV and V.
305

-------
 Phase III,  the  simultaneous  air and  water wash  of  the  coarse  sand
 below fluidization velocities  maintains  the  filter media in cleaner
 condition than  the air  scour and water wash  for.the  dual-media  filter
 and the  auxiliary  surface  wash of the  mixed-media  filter.   Also, it
 would appear the auxiliary surface wash  used on the  mixed-media fil-
 ter resulted in better  backwashing than  the  air scour  and water wash
 as  used  on the  dual-media  filter up  to day no.  180 of  the series,  and
 thereafter they were  comparable.

 The results for Phases  IV  and  V are  shown in Fig.  47.  The  same basic
 patterns similar to Phase  III  exist  here as  well.  The simultaneous-
 air and  water wash of the  coarse sand  below  fluidization velocities
 was superior during the entire period  of Phases IV and V.   During
 the first half  of  operating  period,  up to about day  no.  252,  the
 auxiliary surface  wash  filter  maintained a cleaner filter bed than
 the air  scour,  fluidized water wash  method of backwashing.  After
 that day,  the test results vary a great  deal.   No  pattern exists be-
 tween the dual- and mixed-media filters  as before.   The  only  constant
 aspect of behavior during  this period  was in the coarse  sand  filter
 where very little,  if any, buildup of  solids is shown  by the  test
 results.   However,  the  other two filters and corresponding  method  of
 backwashing show a gradual buildup of  solids on the  media.  Some
 operational difficulties were  experienced throughout this period with
 the surface and subsurface washers,  especially  after day no.  274.
 They either turned very slowly or failed to  turn at  all  at  times.
 This might  explain the  erratic results for the  filter  using surface
 and subsurface  washers.  No  explanation  can  be  given to  qualify the
 results  for the air-scour, water wash  except to say  it was  caused  by
 the simple  failure of the  backwashing  method to maintain the  media
 in  clean condition.   The effectiveness of the combined air-water
 scour on the dual-media filter was no  doubt  hampered by  brief time of
 application,  approximately 30  sec due  to the short rise  distance from
 the media to the overflow  trough.  One can conclude  from the  abrasion
 test results that  the simultaneous air and water wash  below fluidiza-
 tion velocities is  the  superior backwashing  technique.   Also, when
 working  properly,  the surface  and subsurface washer  backwashing aux-
 iliary and  the  air-scour auxiliary as  used in the dual-media  filter
 are  roughly equivalent  in  effectiveness,  with a possible edge in
 favor of  the  former.

 Initial head  loss.   The  initial head loss  readings throughout the
 testing period  were also recorded  in an  attempt to evaluate the ef-
 fectiveness of  the  three backwashing techniques.  The  initial head
 losses were monitored for each day of  filter operation.  All  the
 readings were standardized by  allowing 15  min of filtration before
 reading the piezometer  tubes;  later in the testing period they were
 read  after  the head loss readings  had  stabilized, which  ordinarily
 took  less than  15 min.  Figure 48  graphically displays the  initial
head  loss readings  of all the  filter observation days  and several
 additional data points for non-observation runs around periods of  in-
 terest.
                                 172

-------
U)




k.
C
u-
o
*
CO
o
_J
§•
UJ
X
_l
<
1—
z







2.0
1 1.5
UJ
J 1.0
UJ
X 0.5
5
0.0

1 Q.
. . O \SP
-* 	 . 2.1 gpm/sq ft— *U-~ 3. 2 gpm/sq ft 	 ^ Oo
r» « O
^> ^,0 ° 0 0 <>
~ °b o o ^ o ^ o^ ° ° "
Q cP oo OQ cP 0 0
3 °
^ I 1 1 1 1 1 1 I.I
1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1
O ' 150 170 190 210 230 250 270 290 310
u
                                               CALENDAR DAY, 1974
                     Fig. 48.  Initial head loss data for each filter for entire study.

-------
 Before  one can attempt  to  evaluate  Fig.  48,  the  reader must be  aware
 of several points.   There  were several  changes between Phase III  and
 the rest  of the study.   First  of all, the media  depths and  piezometer
 locations were changed  for the dual-  and mixed-media filters.   There-
 fore,  the points representing  the initial head loss  in Phase III  are
 not under the  exact  same conditions as  Phases IV and V.  For the  dual-
 media  filter,  the initial  head loss represents the head loss across
 the entire bed, 27 in.  that is,  from  the top of  the  filter,  including
 15 in.  of coal, 9 in. of sand, and  3  in. of  supporting gravel.  For
 the mixed-media filter,  the values  represent the head loss  across the
 filter  bed from 3 in. into the bed  to 27 in. into the bed.   Inadver-
 tently, there  was no piezometer tap above the media  bed, so the head
 loss in the top 3 in. (approximately) of coal was excluded  from the
 measurement.   Therefore, the total  head  loss presented corresponds
 to the  head loss across  the 12 in.  of coal,  9 in. of sand,  and  3  in.
 of garnet sand below and excludes the head loss  in die first 3  in. of
 coal at the top of the bed.  Therefore,  valid conclusions or trends
 can not be made for  the  mixed-media filter in Phase  III.  The initial
 head loss for  the coarse sand  filter is  the  head loss across the
 entire  46 in.  of media;  there  was no change  of media or piezometer
 tap location for this filter throughout  the  length of the testing
 period.   At the start of Phase IV,  new bed depths and piezometer  taps
 were established in  the  dual-  and mixed-media filters.  The  initial
 head loss for  the dual-  and  mixed-media  filter now corresponds  to the
 head loss across the total bed depth.  The media used in Phase  III
 was discarded  and new media  of the  same  gradation was installed.
 Also, the filtration rate  was  changed from approximately 2.1 gpm/sq
 ft for  Phase III to  approximately 3.2 gpm/sq ft  for  Phases  IV and V.
 Because of these changes,  and  because of the scatter of the  measured
 values, it is  difficult  to draw  firm conclusions from the initial
 head loss data.

 Several observations can be  made  of Fig.  48.  As stated earlier,  an
 increase  in initial  head loss  is  indicative  of mud balls and agglom-
 erates due to  poor backwashing.   During  Phases IV and V, when valid
 trends can be  identified because  there were  no changes  in experimen-
 tal routine, the mixed-media filter with auxiliary surface  and  sub-
 surface wash exhibits a more pronounced  upward trend,  especially  at
 the  end of the  testing period, than the  other two filters.   The dual-
media filter with air scour  and water wash and the coarse sand  filter
washed below fluidization  velocities show little  if  any buildup in
 initial head loss.  The  surface and subsurface wash   equipment  of the
mixed-media filter experienced operational difficulties in  the  latter
 stages of  Phase V as already discussed.   Since the water fluidization
backwash  alone could not maintain the bed  in clean condition, the
 initial head loss increased.  On day no.  293, the filter bed was
 fluidized  and the surface  and  subsurface washers were  turned manually
by reaching down  into the  bed with a length  of pipe.   The bed was
fluidized  approximately 30 min, and the washers were  turned  around
50  times with the washer supply turned on, in an effort to free the
washer.   In the process, the stirring action broke up  all the mud
                                  174

-------
balls and agglomerates within the bed.  This cleaning of the bed
corresponds to the drop in initial head loss for the next few days.
Since the washers were not freed, the buildup of mud balls returned
shortly and continued to the end of the testing period, and the higher
values of initial head loss returned.  Thus, the rising initial head
loss in the mixed-media filter at the end of the study must be at-
tributed to the non- functioning of the surface and subsurface wash
auxiliary.

Head loss development.  Some typical head loss curves during Phases
III  through V are presented in Figs. 49 to  55.  Only a limited number
are presented because they are all very similar in general appearance.
The curves show the head loss across various depths of media within
the filter bed (with zero depth at the top surface of the media) .
The initial bed depth of each filter was measured before each run and
was used to determine the actual depths of media covering the piezom-
eter taps.  Since the initial bed depths varied throughout the study,
the depths of media reported on the head loss curves varied as well.
All the depths were measured to the nearest 1/4 in. and then rounded
to the closest 0.1 in. for presentation.  Attention is again called
to the inadvertent omission of a piezometer tap above  the media in
Phase III for the mixed-media filter as is evident from the top head
loss curves for that  filter during Phase III.

A casual review of the curves shows the marked advantage of the coarse
sand "filter in terms of run length.  Since all three filters were
operated at identical filtration rates, production would be directly
proportional to the run length.  A comparison between  the filters
with regard to head loss performance can be made in a number of
way?, e.g., run length to a given head loss, volume of production per
unit increase head loss, or solids capture per unit increase in head
loss.  The latter method was selected because it can be used to com-
pare filtration studies conducted at different times and on various
types of influents and is sometimes used to predict head loss develop-
ment in filter design [133].  The influent and effluent suspended
solids concentrations, run length, flow rate, and the  initial and
terminal head loss are the operational data needed to  calculate a
solids capture by the following equation:

Solids capture value   (SS .  f i ~ ss f f j) (Run length) (Flow rate/unit area)

                    -- (Head  loss increase) -
Table 17 presents average solids capture values for all the observa-
tion runs reaching a total head loss of at least 4 ft.  Phase IV is
omitted from the table because of uncertainty about the suspended
solids analysis results and will be discussed later.  As would be
expected from the head loss curves which have been presented, the
coarse sand filter has a higher solids capture value for both series.
The average solids capture values for the dual- and mixed-media fil-
ters are a factor of two or three times smaller than the coarse sand
                                  175

-------
     DUAL MEDIA - AIR SCOUR WASH
     DAY NO. 155
 6 — FLOW RATE 2.15 gpm/sq ft
 5 —
     BED DEPTH
        INITIAL  35.5 in.
        FINAL   34.5 in.
                                0-27.5 in.
   MIXED-MEDIA - SURFACE WASH
   DAY NO.  155
-  FLOW RATE 2.15 gpm/sq ft
   BED DEPTH
      INITIAL 38.8 in.
      FINAL   37.8 in.
  024      6      8      10     120     2     4      6      8      10     12
       TIME FROM BEGINNING OF RUN, hours      TIME FROM BEGINNING OF RUN, hours
   J4
    o
   4:
          COARSE MEDIA-SUBFLUIDIZATION WASH
          DAY NO. 155
          FLOW RATE 2.15 gpm/sq ft
          BED DEPTH
            INITIAL 45.0 in.
            FINAL  45.0 in.
                                 10     12    14     16
                        TIME FROM BEGINNING OF RUN.  hn
             18
                   20
                                                                    0 to 45 in.
                        22
                              24
Fig.  49.   Chronological  head  loss development  at  various media
            depths, day no.  155, Phase  III.
                                     176

-------
  6 -


  5 -




£

031-
        DUAL MEDIA - AIR-SCOUR WASH
        DAY NO. 183
        FLOW RATE 2.05 gpm/iq ft
        BED DEPTH
          INITIAL   35. Jin.
          FINAL   34.3 In.
                                   0 to 27.5 in.
                                 3.5to 27.5 In.
   MIXED MEDIA - SURFACE WASH
   DAY NO. 183  F
_ FLOW RATE 2.05 gpV«l ft
   BED DEPTH
      INITIAL   39,0 in.
      FINAL    	
                                                                              to27in
                                                                            7 to 27 in.
                                                                             1 to 27 In
                                                                           27 In
                                                                           O23to27ln
          2     4     6      8     10
        TIME FROM BEGINNING OF RUN, hn
                                       12
                                                 2     4     6      8     10
                                              TIME FROM BEGINNING OF RUN, hn
                                   12
         COARSE MEDIA - SUBFLUIDIZATION WASH
         DAY NO.  183
       - FLOW RATE 2.07 gpn\Aq ft
         BED DEPTH
            INITIAL 44.0 in.
       _    FINAL   43.3 in.
    §3
                                                         0 to 44 in.
      048
                            16     20     24    28     32
                         TIME FROM BEGINNING OF RUN,  hn
                     40
Fig.  50.   Chronological  head  loss development at various  media
             depths,  day no.  183,  Phase III.
                                         177

-------
    i
    */> -
    03

    Q
          DUAL MEDIA- AIR-SCOUR WASH
          DAY NO. 239
          FLOW RATE 2.92 gpm/sq ft
          BED DEPTH
             INITIAL  38.3 in.
             FINAL   36.5 in.
                    0 to 30.3 in.
      1 —
                    6.3 to 30.3 in.
                    14.3 to 30.3 in.
                     .3 to,30.3 in,|
       MIXED MEDIA - SURFACE WASH
       DAY NO. 239
       FLOW RATE 2.96 gpn/sq ft
       BED DEPTH '
         INITIAL 39.0 in.
         FINAL  37.8 in.
                                                                      Ote27 in.
                                                                      3 to 27 in.
                           7 to 27 in.
                     O—O-o 19 to 27 In.
                           23 to 27 ip.
            2     4      6      8     10
          TIME FROM BEGINNING OF RUN, hn
12  0
         2     4     6     8     10
       TIME FROM BEGINNING OF RUN, hn
12
    .. 5 •
   93
          COARSE MEDIA -SUBFLUIDIZATION WASH
          DAY NO. 239
          FLOW RATE 2.96 gpm/sq ft
          BED DEPTH
             INITIAL  43.0 in.
             FINAL   43.0 in.
                              8     10     12     14     16
                          TIME FROM BEGINNING OF RUN, hn
                                                                  20
                              22
                                                                              24
Fig.  51.    Chronological head loss  development at various media
              depths,  day  no.  239,  Phase  IV.
                                          178

-------
  23
  o
        DUAL MEDIA - AIR-SCOUR WASH
        DAY NO. 240
        FLOW RATE 3. 16 gpm/sq ft
        BED DEPTH
          INITIAL 38. 5 In.
          FINAL  36. 5 in.
                       0 to 30.5 in.
    .5 to 30.5 in.
         30.5 in.
-cyo22.5to30.5in.
                             MIXED MEDIA - SURFACE WASH
                             DAY NO. 240
                             FLOW RATE 3.20 gpm/«q ft
                             BED DEPTH
                                INITIAL  39.5 in.
                                FINAL   37.8 in.
                                                  J) to 27.5 in.
                                                                       3.5 to 27.5 in.
                                                                      7.5 to 27.5 in.
                                                                      ,19.5 to 27.5 in.
                                                                      23.5 to 27.5 In.
          2     4      6     8     10
        TIME FROM BEGINNING OF RUN, hn
                      12   0     2      4      6     8     10
                             TIME FROM BEGINNING OF RUN, hn
12
          COARSE MEDIA - SUBFLUIDIZATION WASH
          DAY NO. 240
          FLOW RATE 3.24 gpm/sq ft
          BED DEPTH
             INITIAL  43 in.
             FINAL   42 in.
                         6     8     10    12     14
                         TIME FROM BEGINNING OF RUN, hn
                                                                          22
                                                                               24
Fig.  52.    Chronological head  loss  development  at  various  media
              depths,  day no.  240,  Phase  V.
                                         179

-------
    8,
    _i 3
                       jOto 29 in.
DUAL MEDIA - AIR-SCOUR WASH
DAY NO. 267
FLOW RATE 3.16 gpm/sq ft
BED DEPTH
   INITIAL 37.0 in.
   FINAL  35.5 in.
                                                               0 to 27 in.
                                                                to 27 in.
MIXED MEDIA - SURFACE WASH
DAY NO. 267
FLOW RATE 3.20 gpm/»q ft
BED DEPTH
   INITIAL 39.0 in.
   FINAL  37.0 in.
                                                               7 to 27 in.
                                                               19 to 27 in.
                                                               23 to 27 in.
            24     6     8     10
           TIME FROM BEGINNING OF RUN, hn
                                          12
                               2     4     6     8     10
                               TIME FROM BEGINNING OF RUN, hn
                    12
       $5-
       I
             COARSE MEDIA - SUBFLUIDIZATION WASH
             DAY NO. 267
             FLOW RATE 3.24 gpm/sq ft
             BED DEPTH
               INITIAL 42.0 in.
               FINAL   	
                           6     8     10    12     14     16

                            TIME FROM BEGINNING OF RUN, hn
                                                 20
                                                       22
                                                            24
Fig.  53.   Chronological  head loss  development  at  various media
             depths,  day no.  267,  Phase  V.
                                          180

-------
                   0 to 28.8 in.
                 DUAL MEDIA - AIR-SCOUR WASH
                 DAY NO.  269
                 FLOW RATE 3.04 gpnvAq ft
                 BED DEPTH
                    INITIAL 36.8 in.
                    FINAL   35.5 in.
                   4.8 to 28.8 in.
                      to 28.8 in.
                   20.8 to 28.8 in.
                                                        0 to 26.3 in.
                                                        2.3 to 26.3 in.
                                   MIXED MEDIA-SURFACE WASH
                                   DAY NO. 269
                                   FLOW RATE 3.08 gpm/sq ft
                                   BED DEPTH
                                      INITIAL  38.3 in.
                                      FINAL   37.3 in.
                                    6.3 to 26.3 in.
                                    14.3 to 26.3 in.
                                    22.3 to 26.3 in.
          2     4     6      8     10
         TIME FROM BEGINNING OF RUN, hn
                    12
24     6      8      10
TIME FROM BEGINNING OF RUN, hn
                                                            12
    I
     S.
          COARSE MEDIA - SUBFLUIDIZATION WASH
          DAY NO. 269
        — FLOW RATE 3.12 gpmAq ft
          BED DEPTH
             INITIAL
             FINAL
42.0 in.
42.0 in.
   0 to 42 in.
                                                     2 to 42 in.
                                                      10 to 42 in.

                                                        to 42 in.
                                8     10     12     14     16
                            TIME FROM BEGINNING OF RUN, .hn
                                          18
                                                20
                                                      22
                              24
Fig.  54.    Chronological  head  loss development  at  various  media
              depths, day no. 269,  Phase V.
                                          181

-------
                          0 to 28.5 In.
                   DUAL MEDIA - AIR-SCOUR WASH
                   DAY NO. 281
                   FLOW RATE 3.08 gpm/tq ft
                   KD DEPTH
                     INITIAL 36.5 in.
                     FINAL  35.0 in.
                          4.5 to 28.5 In.
                          8.5 to 28.5 in.
                          16.5 to 28.5 in.
                                                             0 to 26.5 in.
                                                             2.5 to 26.5 in.
       MIXED MEDIA - SURFACE WASH
       DAY NO. 281
       FLOW RATE 3.08 gprn/m ft
       BED DEPTH
         INITIAL 38.5 In.
         FINAL  36.5 in.
         6.5 to 26.5 in.

         18.5 to 26.5 in.
         22.5 to 26.5 in.
           24      6      8     10
          TIME FROM BEGINNING OF RUN.  hn
                                          12
2     4     6      8     10
TIME FROM BEGINNING OF RUN, hn
                                                                                  12
            COARSE MEDIA - SUBFLUIDIZATION WASH
            DAY NO. 281
            FLOW RATE  3.20 gpm/x, ft
            BED DEPTH
               INITIAL  42.0 in.
               FINAL   41.5 in.
       0 to 42 in.
                           6     8     10     12    14     16
                            TIME FROM BEGINNING OF RUN, hn
                                                                18
                  20
                                                                           22
                              24
Fig.  55.    Chronological head loss development at various media
              depths, day  no.  281,  Phase V.
                                          182

-------
Table 17.  Solids capture per unit head loss results for direct fil-
           tration of trickling filter effluent, 1974.


                                Solids capture value8

Phase III
Phase V
Dual
media
0.081
0.042
Mixed
media
_b
0.038
Coarse
sand
0.156
0.140
 a
 Ib  SS captured/sq  ft/ft head  loss  increase.

 Omitted  because no piezometer in operation above  the media  surface
 during Series  III.

 filter.  This higher value for the coarse sand filter indicates a
 slower rate of head loss development and longer run lengths than for
 the dual- and mixed-media filters.  The dual- and mixed-media filters
 were approximately  the same in Series V, indicating similar head loss
 development patterns.  Therefore, no conclusion about backwashing
 effectiveness between these two filters is warranted from the head
 loss development patterns.

 The shape of the head loss curves also is an interesting finding.
Most of the curves  for the dual? and mixed-media filters show a steep,
 straight increase (i.e., day no. 269) or an exponential curve (i.e.,
 day no. 240).  The  exponential curve is indicative of partial surface
 filtration of compressible solids.  The straightline head loss curve
 found in this study is believed to be caused partly by surface fil-
 tration as well, but bed compression and surface cracking clouded the
 results as discussed in the Phase I results.  However, the head loss
 curves for the coarse sand filter develop at a much slower rate.

An interesting result is shown in Fig. 52 on the head loss curve for
 the coarse sand filter on day no. 240.  The initial head losses were
 recorded, but as the filtration run progressed the head loss or re-
 sistance decreased  gradually then increased gradually.  This pattern
of decrease followed by increase is present in every single head loss
curve for the coarse sand filter.  One possible explanation for this
 strange behavior is the following.  After the air and w.ater wash is
 finished, a water only wash is applied to flush out the dirt parti-
cles within and above the media and to force out the air remaining
 in the bed.  It was observed that even after this water wash there
were a large number of air bubbles still remaining in the bed.  The
bubbles take up volume, which reduces the available space for water
passage, thereby increasing the actual flow rate and initial head
 loss.  It is hypothesized that throughout the filtration run the
 air is dissolving into the wastewater (which is generally below the
                                 183

-------
 saturation level in oxygen)  while the head loss is increasing due  to
 the accumulation of solids.   At the beginning of the run the air is
 dissolving at a faster rate  than head loss buildup which would pro-
 vide an increased filtration area and a lower head loss  value.  This
 occurs up to a point where the oxygen is completely dissolved or the
 rate of dirt accumulation is greater so the head loss development
 curve would then have an increasing pattern..

 Water quality.   The means and standard deviations of the various water
 quality parameters are given in Tables 18 through 20 for Phases III
 through V.   The analytical tests were performed on composite samples
 of the influent and effluents taken throughout each observation run.
 Turbidity measurements were  taken approximately every two hours for
 the entire  length of the filtration run.   All of the values  for a
 particular  run were averaged to arrive at one number representing  a
 turbidity value for each observation run.  The first six tests re-
 ported in the tables,  that is,  BOD,  soluble BOD,  suspended solids,
 TOC, SOC, and turbidity,  represent parameters commonly used  to mea-
 sure the removal efficiency  of  the filters.   For these parameters,
 the average lab test result  value, which appears in Tables 18,  19
 and 20,  was calculated omitting partial data.   If the analytical re-
 sults for any particular observation run were missing a  parameter,
 i.e., BOD,  TOC, etc.,  for one or more  filters,  then all  the  other
 data for that parameter,  and run,  partial data,  were excluded in
 averaging the lab test results.   The remaining parameter averages,
 that is,  NH4,  N03,  N02,  ORG  N,  OP04,  and  TPO^  were calculated  in-
 cluding  all the data obtained from the  lab  tests.

 The same  sampling containers were used  for  Phases  III and IV.   During
 Phase IV, the  lab results  seemed erratic,  so new sample  containers
 were purchased  and more  rigorous washing  of  sample containers was
 commenced at  the beginning with  the  start of Phase V.  It was be-
 lieved  that the old sample containers were  not  being adequately
 washed during Phase IV,  so the  conclusions drawn about filter perfor-
 mance will  be based on Phases III  and V.

 Excluding Phase IV,  the  coarse  sand  filter  effluent proved to be of
 slightly, but consistently,  poorer quality  than the dual- or  mixed-
media filter effluent.  When comparing  the dual-  and mixed-media
 filters, no such  pattern existed.  Comparing just  these  two  filters,
virtually half  of  all  the  effluent quality  tests were  low for each
 filter.  Also,  because of  the small  sample number  and  the large  stan-
dard  deviation,  one would  be  hesitant to  declare  that  even the  coarse
 sand  filter produced the poorest effluent quality.  The  effluent
quality data between the filters are close,  yet  so variable,  that no
firm  conclusions  can be drawn concerning  performance differences.

Filter bed compression.  An  additional observation made  throughout
 the  study was the  amount of compression experienced  throughout  the
 filter run.  Excessive compression is indicative of dirty filter
media caused by  inefficient backwashing procedures.  When dirt
                                 184

-------
Table 18.  Summary of analytical test results for Phase III.
Filter effluent


BOD5 (mg/1)
N = 13a
Soluble BOD5 (mg/1)
N = 10
Suspended solids (mg/1)
N = 14
TOC (mg/1 as C)
N = 12
SOC (mg/1 as C)
N= 12
Turbidity (FTU)
W- 16
Ammonia (mg/1 as N)


Nitrate (mg/1 as N)


Nitrite (mg/1 as N)


Organic nitrogen
(mg/1 as N)

Orthophosphate
(mg/1 as P04)

Total phosphate
(mg/1 as PO^)

« = number of observation
O = standard deviation.

Influent
14.61
0=6. 00b
3.88
0=1.79
37.49
0=12.03
12.51
0=4.53
7.54
0=1.67
16.38
0=4.31
4.62
0=1.77
N«17
5.53
0=1.21
N=14
0.43
0=0.11
N=12
0.70
0=0.42
N=2
10.99
0=2.79
N=13
12.17
0=3.53
N=6
runs averaged,

Dual
media
3.73
0=1.72
1.97
0=0.96
6.84
0=3.23
8.06
0=2.24
6.97
0=1.43
2.38
0=0.97
4.48
0=1.74
N=17
5.92
0=1.50
N=14
0.68
0=0.33
N=12
2.30
0=1.98
N=2
10.18
0=3.06
^=13
9.27
0=2.92
N=5


Mixed
media
4.11
0=2.03
2.20
0=0.99
6.31
0=3.87
8.19
0=1.87
7.21
0=1.40
2.20
0=0.56
4.35
0=1.59
N=14
5.70
0=1.55
N=13
0.74
0=0.39
N=12
2.90
0=2.26
N=2
10.43
0=3.34
N=13
10.60
0=3.82
N=6


Coarse
sand
4.73
0=2.56
2.34
0=1.11
7.92
0=5.80
7.07
0=1.88
7.44
0=2.01
2.89
0=1.10
4.26
0=1.99
N=17
5.17
0=1.41
N-1A
0.65
0=0.34
N=12
4.65
0=5.16
N=2
10.03
0=2.88
N=13
10.49
0=3.03
N=6


                                185

-------
Table 19.  Summary of analytical test results for Phase IV.
Filter effluent

BOD5 (mg/1)
N = 8a
Soluble BOD (mg/1)
N = 8
Suspended solids (mg/1)
N = 8
TOG (mg/1)
N = 8
SOC (mg/1)
N = 8
Turbidity (FTU)
N = 8
Ammonia (mg/1 as N)


Nitrate (mg/1 as N)


Nitrite (mg/1 as N)


Organic nitrogen
(mg/1 as N)

Orthophosphate
(mg/1 as PO.)
^T
Total phosphate
(mg/1 as PO,)
*T
Influent
17.04
i
C7=4. 21b
4.71
0-1.92
36.41
(7=9.22
11.84

-------
Table 20.  Summary of analytical test results for Phase V.
Filter equipment

BOD5 (mg/1)
N = 15a
Soluble BOD5 (mg/1)
N = 15
Suspended solids (mg/1)
N=14
TOC (mg/1)
N = 10
SOC (mg/1)
N = 10
Turbidity (FTU)
N = 15
Ammonia (mg/1 as N)


Nitrate (mg/1 as N)


Nitrite (mg/1 as N)


Organic nitrogen
(mg/1 as N)

Orthophosphate
(mg/1 as PO.)
M*

Total phosphate
(mg/1 as PO,)
M-
Influent
30.38,
a=!4.52b
9.67
0=3.76
34.08
0=16.87
19.86
cr=8.03
13.41
0=3.22
17.60
CT=6.18
21.08
cr=5.30
N=15
3.09
o=i.77
N=6
0.48
0=0.24
N=6
2.13
o=l.28
N=2
24.61
(7=4.17
N=5
23.41
a=0.69
N-=4
Dual
media
12.68
or=6.88
7.21
0=3.72
7.05
0=4.27
12.02
0=4.16
12.00
a=3.98
4.80
CT=2.28
20.00
a=6.24
N=15
2.84
a=2.05
N=6
0.52
a=0.24
N=6
2.17
(7=1.95
N=2
24.67
0=6.09
N-5
20.66
a=3.10
N=4
Mixed
media
12.99
(7=6.82
7.27
(7=3.61
6.82
a=3.io
12.77
(7=3.20
11.83
(7=2.60
6.78
(7=3.01
19.79
(7=5.76
N=15
2.25
(7=1.78
N=6
0.46
(7=0.28
N=6
2.49
0=1.51
N=2
25.04
a=5.05
N^5
21.31
(7=2.69
N=4
Coarse
sand
14.46
a=6.56
7.78
0=3.57
9.46
(7=4.53
12.99
(7=3.96
12.98
(7=3.47
4.66
0-2.12
20.69
(7=5.47
N=15
2.65
0=1.27
N=6
0.44
0=0.15
N=6
1.81
0=0.65
N=2
25.62
0=5.14
N=5
23.39
0=0.75
N=4
  N = number of observation runs averaged.




 O = standard deviation.
                                 187

-------
 adheres to the filter media,  the actual  grains  do  not  contact each
 other but rest upon the dirt  layers that surround  the  grains.  As
 head loss develops during filtration and thus applies  a  load to the
 filter bed, these dirt layers compress,  which results  in a noticeable
 compression of the entire bed.   Clean media show no  compression
 during filtration since the grains  touch one another and no compres-
 sible layers exist.  The head loss  developments curves (Figs. 49
 through 55) indicate the initial and final  depth of  the  entire filter
 bed.  As can readily be seen, the dual and  mixed filters have approxr
 imately 1-in. compression in  the early stages of the study progressing
 up to 2-in. compression in Phase V.   The coarse sand filter shows no
 compression at all early in the  study, with minor  compression later.
 The recorded compression of the  coarse sand filter is  less reliable
 because an uneven surface results from the  combined  air  and water
 wash below fluidization velocity.  Thus,  the depths  recorded are
 average depths by visual estimation and  subject to greater uncer-
 tainty.  These observations of compression  reinforce the conclusion
 previously established that the  coarse sand filter was cleaner than
 the other two filters.   The dual- and mixed-media  filters exhibited
 roughly comparable degrees of compression,  indicating  both were in
 somewhat dirty condition,

 Filter cleanup operations.  After 170 days  or over five months of
 testing,  the three filters were  removed  from normal  service and sub-
 jected to clean up operations.   All  three filters were first subject-
 ed to a normal backwashing  sequence.  This was  followed  by a new
 backwashing technique  to see  if  more dirt could  be removed from the
 media.   Samples of the  dirty  backwash water were collected and com-
 posited during the new  backwash  routine.  The total  solids of the
 three composite samples  were  then determined.

 The cleanup operation consisted  of  a prolonged  air-water combination
 wash below fluidization velocities.   The water wash  rate was 8 gpm/
 sq ft and the air  wash  rate was  4 scfm/sq ft.  The dual-media filter
 was the first filter subjected to this new  sequence.   The filter
 overflow  was  collected  in a 30-gal.  garbage  pail to  catch the media
 being washed  into  the  overflow weir.  The dirty  washwater samples
 were collected at  the weir  overflow  in the  filter.   For  the first 2
 min of backwashing, when samples were taken  every  30 sec, approxi-
 mately 20 ml  of sample was composited.  After the  first  2 min,
 samples were  taken every minute,  approximately 40 ml of  sample was
 composited  for each collection.   This was continued  for  13 min until
 the washwater was  quite  clear in appearance  leaving  the  filter.  Tur-
 bidity measurements were attempted on all the samples, but the water
was  so dirty  it was doubtful  the readings were within  the capabili-
 ties  of the  instrument,  and they are  therefore not reported.  The
 total  solids  of three aliquot samples of the washwater composite was
measured.  The average result of the  three aliquot samples for the
 dual-media  filter was  1019 mg/1.  The underdrain plate was also re-
moved  to  see  if any sand was evident  from leakage  through the dis-
 turbed  gravel.  None was evident.
                                 188

-------
The mixed-media filter was treated to the same cleanup procedure as
the dual-media filter.  The dirt released in the first few minutes
was phenomenal.  The water was literally black with solids and de-
scribed as much dirtier than the dual-media filter.  The media quick-
ly worked its way about 6 in. into the gravel.  The bed was complete-
ly mixed in a few seconds, and as the wash progressed the garnet
became more evident throughout the bed.  Samples were taken and com-
posited in the same manner as in the dual-media filter but continued
for 20 min because it took longer for the washwater to clear up.  The
30-sec sample was diluted 10 to 1 and still recorded a 60 FTU turbid-
ity reading, indicating a 600 FTU condition.  However, since the tur-
bidity readings were so doubtful, the remaining samples were not
measured.  The total solids of the washwater composite for the mixed-
media filter was 1415 mg/1.

The same backwashing procedure and sampling technique was performed
on the coarse media filter except that the air rate was increased to
the usual level for that filter (7 scfm/sq ft).  The backwash was
very similar to those seen throughout the study for that filter and
was continued for only 10 min.  The discharge water was very clear
throughout the wash compared to the dual- and mixed-media filters.
The total solids for the coarse media filter was only 616 mg/1.
Finally, all three filters were drained and air blown through them in
hopes the media could be left through the winter without breaking
the plexiglass window.  No damage was noted the following spring.
Table 21 presents a summary of the data obtained during the filter
cleanup operation.  As indicated by the visual observations, the

Table 21.  Data summary for cleanup operation at the end of the
           operating period in 1974.




Dual media
Mixed media
Coarse sand
Duration of
special backwash
procedure,
min
15
20
10
Total solids of
composite samples
of backwash water,
mg/1
1,019
1,415
616

Total dirt
released,
g/sq ft
1,041
1,928
419
mixed-media  filter  released  the  greatest  total  amount  of dirt during
the  cleanup  operation,  indicating  it was  in  the dirtiest condition
prior to  the cleanup.   The difference between the  filters would have
been more dramatically  evident if  suspended  solids analyses had been
measured  rather  than  the  total solids which  were measured inadver-
tently.   The total  solids values include  soluble inorganic solids es-
timated at about 300  to 400  mg/1 which have  nothing to do with  the
dirt released in the  cleanup operation.   Nevertheless, the cleanup
operation results do  show a  marked advantage for the simultaneous air
                                  189

-------
 and water backwash of  the coarse  sand compared to  the other two fil-
 ters.  The results also  show an advantage of  the air-scour auxiliary
 for the dual-media filter compared  to the surface  and subsurface
 auxiliary of the mixed-media filter.  This  is partly due to release
 of solids from the gravel layers, and from  the inside walls of the
 mixed-media filter upon  the first application of air-water wash, as
 was also noted in Phase  II.

 Other backwash design  investigations.  A number of other observations
 were made in Phases III  through V to try to answer some important
 design questions related to backwashing.  Due to strainer clogging
 problems in Phases I and II, strainers with coarser slots were used
 in Phases III through  V.  These strainers necessitated the use of
 gravel in the underdrains with the  potential problem of gravel move-
 ment, especially when  air and water backwash were used simultaneous-
 ly.  Unfiltered secondary effluent  was used for backwashing.  The
 advantage of this arrangement is that recycled backwash water does
 not increase the hydraulic load to  the filters.  The disadvantages
 are the increased hazard of strainer clogging and  the potential ef-
 fects of the dirtier backwash water on the underdrain gravel.  The
 details of these investigations follow.

 Filter influent "feedwater" was used as a backwash supply for all
 three filters in Phases  III through V.  That supply was the normal
 secondary effluent in  this case.  Two problems arose as a result of
 this backwash supply.

 The first problem arose  after 51 days of operation when there was a
 slight increase in the underdrain pressure  for the dual-media filter.
 Two days later, it exceeded 15 psig (the gage limit) when normally
 it was approximately 3.5 psig.  Only 75% of the backwash rate pre-
 viously used could be  attained even when the control valve was wide
 open.  The next day there was a popping noise, and water streamed
 upward along one side  of the filter during  the backwash cycle.
 Originally, it was thought a nozzle had failed, but closer examina-
 tion showed a gasket around the underdrain plate had failed.

 The media was removed  and the nozzles examined.  A 1/4-in. thick ring
 of coal was found around the inside of the nozzle  slits.  The back-
 plate for the underdrain plenum was removed, and more coal and sand
 was found.  Although no  failure had occurred in the mixed-media fil-
 ter, a similar increase  in underdrain pressure led to the same work
 on that filter.  Coal  and sand were found there as well.  The same
modifications and cleanup were performed on that filter also.  The
 coarse sand filter had no evidence  of strainer plugging at the time
 of this incident.  The low wash rates were believed to be the reason
 for this.  Late in the Phase V, however, some plugging of one strainer
was evident at the window, as reported before.

 It was believed that some media had been washed out of the filter and
pumped back into the underdrain.  Since the influent pump and backwash
                                 190

-------
waste discharge meet in the same collection box, this was a believ-
able explanation.  An apparatus was installed which carried the back-
wash waste discharge out of the collection box, and no similar prob-
lems occurred.

The second problem observed was the accumulation of substantial black
solids in the supporting gravel layers of the dual- and multi-media
filters and in the bottom 15 in. of the coarse sand filter.  It is
not possible to attribute the solids solely to the backwash water or
to the filtered water reaching those depths.  Nevertheless, the gravel
layers were very dirty and gravel movement did occur as described in
more detail in the following paragraphs.

A double reverse graded gravel underdrain (coarse to fine to coarse)
was used in the dual-media filter as previously described to avoid
movement from the combined air and water backwash.  As previously
stated, the double reverse graded underdrain was installed upside
down, and the backwashing with air and water simultaneously caused
almost immediate mounding of the fine gravel.  Excessive mounding
required attention, so the media was removed and the gravel was
placed correctly in the filter two weeks later.  No shifting or mound-
ing of the gravel was noted immediately following or for the next six
weeks following the change until a strainer clogging incident com-
pletely disrupted the bed.  Even larger strainer openings were used
in the filter rebuilding (4.5 ram), so a revised gravel support was
used to accommodate the larger slots as previously described.  No
gravel mounding or shifting was observed during the next three weeks.
However, some temporary cracks (horizontal, 1/2 by 6 in.) were noticed
in the finest gravel layer during backwashing  throughout these weeks.
At about that time, some of the fine middle gravel escaped to the top
of the coarser gravel above.  The details of the escape were not ob-
served or recorded, but it may have happened in a single wash, and
further progressive escape did not seem to occur.  The escaped fine
gravel first appeared on the sides, but later  it was observed to have
moved to the center as well.  During backwashing the fine gravel was
carried up into the silica sand by jets, and some gravel remained
there after fluidization had ceased.

Since the gravel escape and mounding came so suddenly, one would
suspect that the desired water wash rate was exceeded during the com-
bination air-water wash cycle.  The control valve was a 1/4-turn plug
valve, and the flow rate was difficult to control precisely.  The
data book had no evidence of this, however, and no quick jumps or ex-
plosions in ths gravel were noted.

Nevertheless, the potential hazard of gravel movement was demon-
strated, and the opening of horizontal cracks  in the fine gravel dur-
ing backwashing was recorded.  Such cracks  in  a large filter could
result in an upset to the gravel layers.  The  accumulation of  solids
in the gravel no doubt contributed to the pressure drop  across the
                                 191

-------
 fine gravel  during backwashing,  and  thus  to  the  formation  of  the
 cracks.

 A conventional  graded  gravel was used  in  the mixed-media filter.  It
 proved  to  be unstable,  even though washed with water  only  at  normal
 backwash rates  reported as  follows.

 Within  20  days,  the top fine gravel  of the regular graded  support
 gravel  was intermixing with the  coarse support garnet layer.  The
 migration  of the gravel upward and the coarse garnet  downward was
 very pronounced  two days later.   The coarse  garnet progressively
 worked  through  6 in. of gravel while the  fine gravel  broke through
 the  coarse garnet layer and began mixing  with the silica sand.  The
 fine garnet  layer between the silica sand and the coarse garnet layer
 had  already  mixed up into the silica sand in the first filter wash
 and  effectively  disappeared from view.  Finally, on day no. 168,
 after 32 days of operation, the  coarse garnet and the top  fine gravel
 were "totally mixed" and mounding.   The mounding and  shifting caused
 channeling,  leading to more mounding up to the point  when  the filter
 was  rebuilt  due  to the strainer  clogging  incident.

 In the  very  first backwash  after rebuilding, coarse garnet migrated
 into the gravel  base until  just  eight  days later mounding  of  the
 fine gravel  and  extensive mixing of  the coarse garnet and  gravel was
 noted.  Once again,  the gravel mounded on the sides as before causing
 jetting action.   On day no. 234  the  course garnet had worked  to with-
 in 3 in. of  the  false  bottom.  The same reports  of coarse  garnet and
 fine gravel  migration were  recorded  to the end of the study.  Unlike
 the  double reverse graded gravel, the  underdrain problems  encountered
 with the regular gravel were well documented and occurred  gradually.

 One  can draw several wastewater  filter design conclusions  from the
 foregoing  observations.   First,  there  are inherent dangers in the use
 of feedwater for backwashing which must be recognized.  For example,
 if gravel  is used in the underdrain, and  if  the  backwash water should
 accidentally contain an unusually high concentration  of suspended
 solids, these solids may be partially  removed in the  fine,  gravel.
 This could result in sufficient  pressure  drop across  the gravel layer
 to lift the  gravel and  cause an  upset  of  the gravel.

 Second, the  double reverse  graded gravel  design  used  in the research
 did  not prove adequately stable  to resist movement when backwashed
with air and feedwater  simultaneously.  The  solids in the  backwash
water may have contributed  to the instability, and the research
 should be  repeated using filtered water for  a backwash supply.

Third, the conventional  graded gravel  support was unstable when
washed with  feedwater alone.  This observation reiterates  the weak-
ness of the  conventional  design  reported  by prior workers  [10,11].
                                 192

-------
Therefore, the use of underdrain strainers without supporting gravel
is a desirable design arrangement, but the use of fine slots of less
than 1 mm should be avoided, and feedwater should not be used as a
backwash.  The advantages of using feedwater do not justify the risks
which result therefrom.

Summary and Conclusions - Phases III through V

The objectives of the experimental investigations of'these phases
were to compare the effectiveness of three backwashing techniques on
three different types of filters while filtering secondary effluent,
and to look at the problems of underdrain strainers and supporting
gravel when backwashed with unfiltered secondary effluent.  The fil-
ters and backwashing techniques were as follows:

1.   A dual-media filter backwashed with air scour followed by water
     fluidization backwash.  The filter media was supported on a
     double reverse graded gravel.

2.   A mixed  (triple) media backwashed with a surface and subsurface
     wash auxiliary before and during water fluidization backwash.
     The filter media was supported on a conventionally graded
     gravel.

3.   A coarse  sand media with a deeper bed backwashed with  air  scour
     and water simultaneously at subfiLuidization velocity.  This  fil-
     ter was  supported directly on the underdrain strainers without
     the use  of gravel.

The experimental data were collected over a five-month period of  con-
tinuous operation of the filters at the Ames, Iowa, trickling filter
plant, using  pilot-scale equipment.  All three  filters were back-
washed with unfiltered secondary effluent  (i.e., feedwater) through-
out the course of the  study.  The following conclusions resulted  from
the study.

1.   Of  the three backwash methods, simultaneous air  scour  and  sub-
     fluidization water backwash of the coarse  sand media proved  to
     maintain the cleanest  filter media based on the  abrasion test
     results,  bed compression data, visual observations,  and a  termi-
     nal  cleanup operation.

2.   The  use  of  air  scour  and subfluidization water backwash simul-
     taneously on coarse  sand media was  able  to keep  the  filter bed
     completely  free of mud balls, but  there were  dirty regions in
     the bottom  15  in. of  the filter  and  along  the vertical corners
     of  the bed.  These dirty regions did  not  impair  the functioning
     of  the  filter,  and  those in  the  corners would be inconsequential
     in  full-scale  filters.
                                  193

-------
3.   The surface and subsurface backwash auxiliary in the mixed-media
     filter and the air-scour auxiliary in the dual-media filter pro-
     vided equivalent backwashing effectiveness.  Both filters ex-
     perienced mud ball problems and other indications of dirty fil-
     ter media.  Comparing these two methods alone, the air-scoured
     filter was cleaner based on terminal cleanup observations,
     whereas the filter with surface and subsurface auxiliary was
     cleaner based on visual observations of the media and slightly
     lower abrasion test results.

4.   Both the double reverse graded gravel support and the conven-
     tional graded gravel proved unstable as used in this research.
     The work should be repeated using filtered water as a backwash
     supply before deciding on the suitability of the use of gravel
     in wastewater filters.

5.   Feedwater is not recommended as a backwash water source because
     of the danger of clogging underdrain strainers and/or gravel.
     The advantages of using feedwater do not justify the risks which
     result therefrom.

6.   The underdrain orifice or strainer system should have suffi-
     ciently large openings so that solids in the backwash water do
     not cause progressive clogging problems.  Media-retaining
     strainers with slots less than 1 mm are not recommended.  This
     recommendation dictates the use of a sufficiently coarse filter
     media or supporting gravel to prevent loss of media to the
     underdrains.

7.   Comparing the three filters of this study, the coarse sand fil-
     ter produced a filtrate slightly poorer in quality than that
     produced by the dual- and mixed-media filters but provided sub-
     stantially more filtrate to a common terminal head loss.  These
     differences can not be attributed to the backwashing methods
     used, but rather to the differences in the filter media.

                   Operation and Results - Phase VI

             Coarse Sand Filtration of Secondary Effluent

This final research phase was devoted to further studies of coarse
sand filters backwashed with air and water simultaneously at subflu-
idization velocity.  The favorable results with one filter of this
type in Phases III through V prompted this final phase to answer some
additional questions about such filters.  It was concluded in Phases
III through V that this type of filter and backwashing routine re-
sulted in a cleaner filter bed than the dual- and multi-media beds
which were studied.  However, there was a layer of dirty sand ob-
served in the bottom of the filter about 15 in. deep.  It was also
observed that the filtrate was slightly poorer from the coarse sand
filter than from the dual- and multi-media filters.
                                 194

-------
Therefore, Phase VI was conducted to determine (1) the effect of
depth of media on the extent of the dirty zone at the bottom of the
filter and (2) the effect of depth of media on the filtrate quality.

To study these questions, the three pilot filters were equipped with
new filter media of slightly coarser size, with depths of 24, 47, and
60 in.  The shallowest filter media was supported on a double reverse
graded gravel of revised design to gain further experience on the
stability of the gravel when backwashed with air  and water simultane-
ously.  The three filters were operated for two months in 1975 fil-
tering secondary effluent at the Ames, Iowa, trickling filter plant.
The secondary effluent was used for backwashing throughout Phase VI.

Operation - Phase VI

Operating routine and sampling details were identical with Phases III
through V.  The filters were backwashed every 24 hr.  Observation
runs were conducted twice each week, with detailed observations re-
corded and composite samples collected for chemical analysis.  De-
tails of media size, gravel gradations, and underdrain strainers have
been described previously in the equipment section.

Backwashing.  The backwashing routine for the first 35 runs of Phase
VI included the simultaneous use of air at 7 scfm/sq ft and water at
8 gpm/sq ft for 15 min followed by water  alone at 8 gpm/sq ft  for 3 min,
This was a slightly lower water flow rate than used in Phases  III
through V and was selected because this is the normal recommendation
of one filter manufacturer who promotes this type of filter in the
United States (Dravo Corporation).  It was visually evident from the
start that the three filters were not as clean after backwashing in
Phase VI as the single coarse sand filter had been in Phases III
through V.  Therefore, beginning with run 36 of Phase VI, the  back-
wash flow rate was increased to 15 gpm/sq ft during both  the combina-
tion air-water wash and  the water wash (alone) which followed.  The
period of combination wash was reduced to 10 min  to maintain the
total water usage for each backwash approximately unchanged.   This
routine continued to the end of Phase VI  at run 51.

Results - Phase VI

The results of Phase VI will not be presented in  as much  detail  as
were prior phases because the objectives were more limited.

Visual observations.  The condition of the  filter media was observed
during and after each backwash.  Prior to run 35  (lower rate of  water
backwash) the condition  of the three  filters was  roughly  comparable.
The descriptions of the media as reviewed through the plastic  window
after the backwashes had been completed are characterized by the
following comments.  The top 6 to  18  in.  of the media appeared clean
except for dirty strips  along the  vertical  corners of the filters.
These strips varied in width from  2 to 5  in.  Below  the clean  area of
                                  195

-------
 the media,  the sand was  progressively dirtier toward  the  bottom of
 the filter.   At times, the  clean region was  not  continuous with,  for
 example,  a  6-in.  clean area at the surface and another  similar  area a
 bit deeper,  separated by a  horizontal dirty  region  in the bed.   Small
 black anaerobic strips and  spots were generally  evident but changed
 position  with time  in the dirtier regions of the bed.   No mud balls
 were observed on  the surface or anywhere within  the bed.

 The air-water backwash action was described  as a feeble pulsing of
 the top 10  to 15  in. of  the bed,  which was most  noticeable in the
 first couple  of minutes  of  the air-water backwash.

 Due to the  relatively dirty condition of the bed reported above,  the
 water backwash rate was  increased beginning  with run  35,  as previous-
 ly  described.   The  condition of the media improved  immediately  and
 continued in  a steady good  condition  for the remainder  of the study.
 The descriptions  of the  media for all three  filters at  the completion
 of  each backwash  are similar.   The media was consistently clean to
 within 12 to  15 in.  of the  filter bottom except  for dirty strips
 along the vertical  corners  of  the filter.  These strips varied  from
 0 to 2 in.  in width.  The bottom 12 in. was  quite dirty except  for
 two clean penetrations immediately above the two underdrain strainers
 closest to  the window.   One or both of these penetrations would reach
 nearly to the strainer level at the bottom.

 The north filter with supporting  gravel was  slightly different.  The
 gravel remained dirty for its  full 12-in. depth.  The clean penetra-
 tions  above the strainers would reach the gravel surface, but a dirty
 region between them would reach 6 in.  above  the  gravel.

 Small  black anaerobic regions were occasionally  reported  in the dirty
 strips in the  vertical corners  of the  filter.  They were  seldom ob-
 served in the  bottom 12-in.  dirty region of  any  of  the  three filters.

 The air-water  backwash action was described  as a good pulsing action
 throughout the bed  except for  the bottom 12  in.  and in  the vertical
 corners.  These regions  remained  dirty as a  result.

 In  view of the marked difference  observed before and after run  35,
when the  rate  change was made,  some detailed observations of the
media  movement were made at  the end of Phase VI.  The water rate was
 increased stepwise  from 8 to 20 gpm/sq ft.   The media action was
 carefully recorded.

The  general behavior was an  increase  in the  vigor and extent of
pulsing as the water rate was increased.  In addition,  and more im-
portantly, it was noted that a  circulation of the sand  occurred and
that the rate  and the circulation increased with the water flow rate.
Apparently,  the rising air and water moved sand up  in the center of
the bed,  because the sand was observed to move down slowly at the
                                 196

-------
window.  The observed details at different rates are summarized in
Table 22.

Table 22.  Action of simultaneous air and water backwash on coarse
           sand at subfluidization velocities,a  (Air rate =
           8 scfm/sq ft for all observations.)


Water rate,
 gpm/sq ft                    Action of media


     8           Slight pulsing on right side in top 3 ft.  No
                 downward circulation patterns.

     10           A bit more vigorous pulsing in a little wider zone
                 on right side in the top 4 ft of the bed.  No
                 downward circulation patterns.

     12           Pulsing full width of bed in top 1 ft, in left
                 side of top 3 ft and in right side of top 4.5 ft.
                 Very slow downward media circulation in the center
                 at about 0.5 in/min.

     15           Pulsing about same as at 12 gpm/sq ft.  Downward
                 circulation of media over entire width (except for
                 2-in. wide strips at corners) at a rate of 2 to
                 3 in./min.  Movement downward continues to 18 in.
                 from the bottom.

     20           Similar but faster action with downward circula-
                 tion of media at 4 to 10 in./min., the lower rate
                 observed where the rising pulsations are more pro-
                 nounced.


aAs  observed in a 1.5 by 1.5 ft filter with 5 ft of sand and 2.5 to
 3.7-ram size, water temp = 21 °C.   No fluidization of the  bed was
 observed at any of the flow rates in the table.

It is believed that this circulation of media is essential for good
cleaning because it moves the media periodically through regions of
intense upward pulsing and movement in which the better cleaning ac-
tion occurs.

At the completion of the observations reported in Table 22, the sand
was  removed from the filters, with special attention to the condi-
tions in the bottom dirty region of the bed.  It was noted that cir-
cular zones of clean media were found directly above the five  strain-
ers  in the dirty region of the bed.  In the north filter with  gravel
support, the above pattern was noted from about 3 to 10 in. above the
gravel.  In the west filter, without gravel, it was noted about 12 in.
                                  197

-------
 from the bottom of  the  filter.  No details were  reported  for  the
 south filter.   It was difficult to tell  the upper  extent  of the zone
 because part of the sand  fell out of the filter  spontaneously when
 the back plate  of the filter box was removed  to  facilitate removal of
 the sand.

 It was also noted that  some filter sand  had escaped  into  the  under-
 drain plenum in the west  and south filters, which  were not equipped
 with gravel.  About two pounds was reported in the west filter.  This
 observation is  not  surprising since  these filters  had strainers with
 4.5-mm slots, and the media was 2.5  to 3.7 mm in size range.  In
 fact,  it is surprising  that more media did not move  downward  through
 the strainers.  The use of the 4.5-nm strainer for this sand  was an
 oversight; it should not  have been used.

 Underdrain gravel stability.  The double  reverse graded gravel in the
 north filter again  proved to be not  completely stable.  On about 407«
 of the backwashes,  a horizontal crack was observed to open in the
 finest gravel layer at  the beginning of  the air-water backwash.  The
 thickness of the crack  varied from 1/8 to 3/4 in.  and the length from
 a few inches to the full  width of the filter.  It  generally contract-
 ed as  the wash  proceeded.

 The gravel did  not  upset  in Phase VI as  it had in  Phase IV, but the
 potential for upset remains evident  in these  observations.  There-
 fore,  the prior conclusions drawn from Phases III  through V about the
 use of gravel in filters  washed with air  and  feedwater simultaneously
 remain unchanged.

 Water  quality.  Fewer water quality  parameters were  measured  in Phase
 VI than in previous phases due to budget  limitations.  The means and
 standard deviations for the analyses of  the composite samples are
 reported in Table 23.   There was no  apparent  difference between the
 performance of  the  three  filters of  different depth.  There appears
 to be  a very slight advantage in the 60-in. depth  as measured by sus-
 pended solids and turbidity, but it  would be  impossible to prove it
 statistically.

 By comparing the values of each run  using a ranked analysis,  some re-
 inforcement of  that conclusion is obtained.   In  this analysis, the
 rank of 1 is assigned to  the lowest  value, the rank  of 2  to the mid-
 dle value, and  the  rank of 3 to the  highest value  for a particular
 filter run.  Comparing  the turbidity values in this  fashion,  the 60-
 in. deep filter had seven lowest, three middle,  and  one highest
 value  for a total score of 16.  The  24-in. filter  had a total score
 of  22, and the  47-in. filter a total score of 28.  The 24-in. filter
 was  supported on 12 in. of gravel in a rather dirty  condition, which
 may have contributed to the filtration and resulted  in its exceeding
 the performance of  the  47-in. filter.  Using  a ranked analysis in the
manner above also placed  the 60-in.  filter in first position  in sus-
 pended solids removed and BOD5 removed.
                                  198

-------
Table 23.  Results of analyses during direct filtration of secondary
           effluent (Phase VI) from June 24 through August 2,  1975,
           using course sand filters of different depths.


                                             Filter effluent

Suspended solids (mg/1)
a (N = il)c
Turbidity (FTU)
a (N = ll)
BOD5 (mg/1)
a (N - ll)
Soluble BOD5 (mg/1)
a (N = ll)
Filter
influent
31.3
9.7
12.6
3.14
15.6
4.7
5.3
1.7
N filter,
24-in.
depthb
5.9
2.1
3.30
1.21
6.5
2.8
3.9
1.3
W filter,
47-in.
depth
6.4
2.3
3.38
1.14
7.1
2.5
4.0
1.3
S filter,
60-in.
depth
5.7
1.8
3.14
1.14
6.6
2.5
3.8
1.3
 Filtration rate 3.0 gpm/sq ft.

 Filtrate from 24 in. of sand and 12 in. of supporting gravel.
^
 CT = standard deviation, N = number of observations.

Looking at the mean values in Table 23, one must conclude, however,
that there was little gained by using the deeper media.  In the fil-
tration of secondary effluent, one would need to use alternate ap-
proaches to reach a higher quality filtrate, either chemical pre-
treatment or a finer filter media.  Deeper media should prolong the
period of acceptable filtrate if higher filtration rates and/or ter-
minal head losses had been used.  In this work, however, deteriora-
tion of the effluent at the end of the filter run was not observed
on any of the filters.

Head loss patterns and initial head loss.  The shape of the head loss
development curves in Phase VI were very similar to those observed
for the coarse sand filter in Phases III through V.  Therefore, no
additional curves are presented for Phase VI.  Furthermore, there was
no apparent difference between the curves for the three filters of
different depth in Phase VI, except for a different initial head
loss.  As expected, the initial head loss was higher for the deepest
filter since they were all operated at the  same filtration rate.

An analysis of the head loss developed to a common  time for each ob-
servation run is summarized in Table 24.  A total of 11 observation
runs are included, the same 11 runs for which water quality data were
                                  199

-------
Table  24.   Mean total head loss during filtration of secondary
            effluent on coarse sand filters during Phase  VI.
N filter,
24 in.
sanda
Mean total head loss , ft 4.49
CT 0.48
W filter,
47 in.
sand
4.35
0.55
S filter,
60 in.
sand
4.54
0.49
&Head  loss  includes  12  in.  supporting  gravel.

 Head  loss  for  an  average run  length of  16.18 hours  (o =  1.89 hours).

presented in Table 23.  It  is apparent that very little difference in
head loss was observed between the three filters.  This would be ex-
pected since they have the  same media size and the suspended solids
removed was nearly the same for three filters, as shown previously.

The initial head loss for the three filters was different due to the
different depths of media.  One observation of interest is the change
in initial head loss that occurred when the backwash flow rate was
increased after run 35.  The average initial head loss values for a
few filter runs before and  after the change in backwash rate are pre-
sented in Table 25.

Table 25.  Average initial head loss for three coarse sand filters
           before and after run 35 in Phase VI, when increase of
           backwash rate was adopted.
Period
Runs 26-34
Runs 36-42
Runs 43-47
(inclusive)
(inclusive)
(inclusive)
N filter,
24 in.
sand*
0.84b
0=0.07
0.45
0=0.03
0.64
0=0.04
W-filter,
47 in.
sand
0.89
0=0.04
0.55
O=0.00
0.64
0=0.07
S filter,
60 in.
sand
1.15
O=0. 09
0.69
O=0. 03
0.82
0-0.05
      loss includes 12 in. of supporting gravel.

 All values in feet.
                                 200

-------
It is clearly evident from Table 25 that a marked reduction in ini-
tial head loss occurred at the higher backwash rate.   This indicates
that the bed was somewhat dirty before the change and that the new
backwash routine achieved a cleaner bed.

There was again some deterioration indicated by the initial head loss
in the later runs of Phase VI, but the series was not continued long
enough to see if that trend continued.

It would be desirable to compare these initial head loss data with
values for the clean media at the beginning of Phase VI.  Unfortu-
nately, the operating routine used prior to Run 26 was not such as to
obtain reliable initial head loss data.  This weakness did not in-
fluence the head loss data for the bulk of each run, only the initial
value recorded.

The mean removal of suspended solids, the mean run length and in-
crease in head loss (total head loss - initial) and the filtration
rate were used to calculate the solids  capture value  as explained  in
the Phase  III through V discussion.  The average value for all these
filters in Phase VI is 0.16 Ib SS captured/sq ft/ft of head loss in-
crease.  This value compares favorably with the value obtained in
Phases III and V in Table 17.

Media loss in air-water backwashing.  The primary advantages of
coarse sand  filters washed with air  and water are (1) more effective
backwash and  (2) lower headloss development, and thus longer filter
runs.  The primary danger of this  type  of filter and  backwash  is  the
potential loss of media which can  occur during overflow due to the
violence of  the air-water action.

Some preliminary observations on this question were made  in the lab-
oratory using a 5.5-in. diameter filter column of transparent plastic
construction.  Sands  of  two sizes  were  observed  in the column while
being washed with air and water simultaneously at various rates.   The
sand depth was 24 in.  The height  to which  the sand  grains were
thrown by the backwash was observed  visually.  This  was admittedly
difficult to judge because of  the  presence  of the air bubbles  and the
violence of  the motion.

The  results  of  the observation are summarized in Table  26 for  the two
sand  sizes.  The data are  admittedly very limited and should  be re-
peated covering  a broader  range of flow rates and media sizes.  They
are  presented here  for the record, merely to  call attention to the
danger  of media loss, and  the  need for  adequate  distance  from the
media to the overflow in the  filter design.

No loss  of  sand was  observed  during the two months  of observation of
 the  three  filters  in Phase VI.  No loss was expected because of the
 large distance  which existed  from bed surface to overflow, a minimum
 of 3 ft in the  filter with 5  ft of media.
                                  201

-------
Table  26.  Height  that  sand  is  thrown by simultaneous  air  and water
           backwash.
                                      Height  sand  thrown
water rate,
gpm/sq ft
2
2
4
4
6
6
Air rate,
scfm/sq ft
4
8
4
8
4
8
1 to 2- mm sand, in.
11-12
12-13
11-12
13-14
15-16
16-17
2 to 3 . 6-mm sand , in .
nil
nil
5-6
5-6
5-6
5-6
Some small loss of media appeared to have occurred in prior Phases
III through V on the single course sand filter then in use.  During
five months of operation, the recorded depths of filter media de-
creased 2 to 3 in.  The uneven surface of the sand made the surface
depth measurement difficult, and thus the loss reported above is only
approximate.  The distance from bed surface to overflow in Phases III
through V was only about 12 in.  In view of this reported loss, a
distance of more than 12 in. should be provided with 2 to 3.6-mm sand
to prevent loss.

Summary and Conclusions - Phase VI

The objectives of the experimental investigation of Phase VI were to
observe the effect of depth of media on the backwashing effectiveness
of deep coarse sand filters washed with air and water simultaneously
and to observe the effect of depth on filtrate quality.  Three pilot
filters were operated in parallel while filtering secondary effluent
at the Ames, Iowa, trickling filter plant.  The filters were equipped
with 24 in., 47 in., and 60 in. of coarse sand in a size range of 2.5
to 3.7 mm.  Double reverse graded gravel was used to support the sand
in the filter with 24 in. of sand.  All three filters were backwashed
with secondary effluent (i.e., feedwater) throughout the two-month
period of operation.

The following conclusions have resulted from the investigations in
Phase VI:

1.   A backwash routine including simultaneous use of air at 7 scfm/
     sq ft and water at 8 gpm/sq ft for 15 min was not effective in
     keeping the filter sand in clean condition.

2.   A backwash routine including the simultaneous use of air at
     7 scfm/sq ft and water at 15 gpm/sq ft for 10 min was effective
     in keeping the filter sand in clean condition except for about
                                  202

-------
     a 12-in. layer at the bottom of the filters.  The extent of
     this dirty layer at the bottom was independent of the depth of
     media and probably is a function of the underdrain strainer
     type and spacing.

3.   In backwashing of coarse sand filters with air and water simul-
     taneously, the rates of flow of air and water should be selected
     to ensure that a modest circulation of the sand occurs in the
     filter, upward above the underdrain strainers and downward be-
     tween the strainers.

4.   The filtrate quality produced by the three filters of different
     depth was nearly the same when filtering secondary effluent.
     Thus, one cannot achieve much improvement in filtrate quality
     with such filters and feedwater merely by increasing the depth
     of media.

5.   The double reverse graded gravel backwashed with feedwater
     proved somewhat unstable again in Phase VI, and the prior con-
     clusions about the suitability of feedwater and gravel are thus
     unchanged as a result of the Phase VI experiments.

6.   The principal advantages of coarse sand filters washed with air
     and water simultaneously were again demonstrated in this phase,
     namely, lower head loss development and thus greater solids
     capture per unit head loss development, and better backwash ef-
     fectiveness.  The principal hazard in this backwash arrangement
     is the potential loss of filter media during backwash overflow
     due to the violence of the air-water action.  Adequate freeboard
     must be provided and the rates of air and water flow must be
     selected to ensure no loss of media.
                                  203

-------
          IX.  EXPANSION AND  INTERMIXING  OF MULTI-MEDIA FILTERS


                              Introduction

 Granular  filters  are widely  used  for water treatment  and  are  gaining
 importance  for  tertiary treatment of wastewater  because of  the recent
 higher effluent quality standards.  In a conventional single-media
 rapid  sand  filter,  the  sand  particles are hydraulically graded during
 backwashing,  resulting  in  the finest particles being  in the upper
 layer  .   This stratification remains after backwashing.   In a strati-
 fied filter bed,  the pore  openings between the particles  vary direct-
 ly with the particle size.   Because of this, most of  the  material
 removed by  the  filter during filtration  is at or near the surface of
 the filter.   An ideal filter would have  this stratification reversed
 so that the pore  size would  be  the largest in the top layer and
 steadily  decrease in size  to the  bottom  layer of the  filter.  The
 recent development  of multi-media filters with media  of anthracite
 coal,  silica  sand,  and  garnet sand approach this ideal pore size
 arrangement.  The densities  of  these three media vary with the an-
 thracite  coal having the lowest density, the garnet sand  the  highest
 density,  and  the  silica sand between the two extremes.  The average
 particle  sizes  of the anthracite  coal, silica sand, and garnet sands
 are selected  to decrease in  that  respective order.  With  proper se-
 lection of  particle sizes, the  filter will stratify with  the  anthra-
 cite coal on  the  top, the  garnet  sand on the bottom,  and  the  silica
 sand in the middle.  Although the pore size may  increase  with depth
 within each individual  media, the overall effect will be  a decrease
 in pore size  of the filter with increasing depth.  This decrease in
 overall pore  size with  increased  depth will greatly increase  the
 penetration of  filterable  solids  into the filter bed, thereby in-
 creasing  utilization of the  filter bed and extending  the  filter run
 length, hopefully without detriment to the filtrate quality.

 With the  addition of anthracite coal and garnet  sand  in dual  and
 multi-media filters, new problems have arisen.   What  degree of expan-
 sion of the individual  media will provide optimum cleaning during
 backwashing?  Is  there  an optimum degree of intermixing at the inter-
 face of the individual  media?  Can the degree of intermixing  of the
media be  predicted and  controlled through selection of backwashing
 rate and media  size?

The specific  aims of this research are fourfold:

 1.   to evaluate  the effectiveness of existing expansion models for
     predicting the expansion of  garnet sand, silica  sand, and coal,
     and  to suggest new or modified models if necessary,

2.   to test  the validity and sensitivity of two available models for
     the prediction of  intermixing between silica and garnet  sands,
     and between coal and silica  sands,
                                 204

-------
3.   to observe the fixed-bed hydraulic profiles of various coal and
     sand filters selected to produce different amounts of intermix-
     ing in order to study the effect of the intermixing on the per-
     meability of the intermixed zone of the filter, and

4.   to observe the effect of interfacial intermixing on the perfor-
     mance of dual-media filters.

           Dual-Media and Multi-Media Filtration Literature

Media Design Characteristics

Baylis et al. [3] gave these typical dual-media design characteris-
tics.  Sand medium usually has an effective size (ES) of 0.40 to 0.55
mm and a uniformity coefficient (UC) of 1.3 to 1.7.  The sand medium
should be clean and well graded and have a specific gravity greater
than 2.65.  The depth of the sand bed is normally 6 to 12 in.  The
top 1/2 in. or more of sand is usually removed after hydraulic grad-
ing to prevent the head loss at the surface from being excessive.
Similarly, the anthracite coal layer usually has an effective size
of 0.8 to 1.2 mm, a uniformity coefficient of 1.3 to 1.7, a specific
gravity greater than 1.5, and a hardness factor on the MOH scale be-
tween 2.0 and 3.5.  The coal medium should be clean and free of all
thin or scaly pieces, often prevalent in smaller sized coal parti-
cles.  Depth of the coal medium is partially dependent on the uni-
formity coefficient.

According to Camp [24], the top layer of anthracite coal has inter-
stitial spaces approximately 20% greater in volume than those of the
top sand layer.  The larger void capacity of coal absorbed more sol-
ids per volume of filter medium.  As a consequence, Camp  [24] suc-
cessfully used filtration rates up to 6 gpm/sq ft and  achieved longer
filtration runs.  Walker [137] suggested a dual-media  filter is best
designed by first choosing a favorable coal size for filtering the
influent water.  The sand size should be chosen to complement the
coal size.

Baylis et al.  [3] considered that the coal  layer should contain the
bulk of the filtering capacity.  The sand layer was then  used to fur-
ther polish the water after  it had passed the coal  layer.  Coal par-
ticles should not be present at  the bottom  of the  sand layer since
larger pore spaces  occur around  the coal particles.  These large pore
spaces lessen  the effectiveness  of the  sand layer  as a good  filter.

Observed Intermixing in Dual-Media Filters

Conley [32] found that the  intermixing  of coal  and  sand  avoided  the
rapid buildup  of head loss  at  the interface, while  acceptable water
quality was  still achieved.  Camp [23]  disagreed with  Conley [32]  by
recommending  a non-intermixed  filter and  suggesting that the good
water  quality results obtained by Conley were due  to excellent
                                 205

-------
 chemical  control.  According  to  Camp  [23],  removal  of  suspended  sol-
 ids was more  efficient  if  the fine  sand  was not  intermixed with  the
 coarse anthracite  coal.

 Later, Robeck and  Kreissl[103] studied the  effect of intermixing in
 coal  and  sand filters at Erie, Pennsylvania.   They  considered  inter-
 mixing beneficial  for increasing the  run length.  They demonstrated
 that  run  lengths increased with  increasing  size  of  the surface coal.
 This  was  achieved  by removing the coal finer than 1.0  mm  from  the
 source medium.

 The amount  of intermixing  at  the interface  was studied by using  hy-
 draulic profiles of  three  different graded-media filters.  The dual-
 media filter  contained  18  in.  of 1.14-mm ES coal and 6 in. of  0.43-mm
 ES sand,  while  the two, single-media  filters  contained 24 in.  of
 0.75-mm ES  coal and  24-in. of 0.43-mm ES sand, respectively.  The
 coal  and  sand media  were unskimmed, and  both  created a large head
 loss  at the surface.  Limited data concerning the hydraulic profiles
 through the three  filter beds were presented  for downflow with clean
 water.  Cumulative head losses were 1.83 ft for  the unskimmed  coal,
 4.75  ft for the unskimmed  sand,  and 2.50 ft for  the dual-media filter
 at 14.5 gpm/sq  ft.   The intermixing zone depth could be determined
 since the hydraulic  profiles  gave evidence  that  the head  loss per
 unit  depth  gradually increased from the  coal  to  sand layer through
 a 6-in. bed depth.   Robeck and Kreissl [103]  also showed  that perme-
 ability in  the  upper sand  layer  can be controlled by varying either
 media to  produce different amounts of intermixing or by varying  the
 backwashing shutdown procedure.   An instantaneous shutdown shifted
 the level of  maximum head  loss upwards 2 in.  from the  level of maxi-
 mum head  loss for  the slow shutdown.  In addition, the instantaneous
 shutdown  gave a greater maximum  head  loss.  Thus, the  instantaneous
 shutdown  gave less available  capacity.

 Brosman and Malina [19] studied  intermixing of dual-media filters
 made  from 18  in. of  0.43-mm ES silica sand  having a UC of 1.32 and a
 specific  gravity of  2.64;  and  6  in. of 1.00-mm ES anthracite coal
 having a  UC of  1.40  and a  specific gravity  of  1.77.  Four of the many
 conclusions were:  (1) intermixing gave  more uniform bed porosity
with depth; (2) increased  intermixing decreased head loss at the  in-
 terface;   (3)  increased initial downflow  head  loss accompanied in-
creased backwash rate and  shorter backwash  valve closure  time for
 intermixed  filters;  and (4) greater size ratios of anthracite to  sand
gave greater  intermixing.  The intermixing  zone was defined as the
 filter length which has, within  each  infinitesimal section, quanti-
 ties of media equalling at least  20%  of  the dry weight of each medi-
um.   Intermixing was not considered to have occurred if the inter-
mixing zone was less than 5% of  the total bed height since flow pat-
terns and bed instabilities probably  caused any observed  intermixing.

Brosman and Malina [19], in studying  filter performance, concluded
that intermixing in dual-media filters resulted in longer filter
                                 206

-------
runs, more uniform head loss with depth, and better filtrate quality.
However, in the opinion of the authors of the present report, this
conclusion may not be valid because the initial cumulative head
losses in the various dual-media filters were unequal; thus valid
comparisons of filter performance were impossible.

                    Backwashing of Granular Filters

Since the advent of the rapid sand filters around the turn of the
century, the study of hydraulics of the filtration process has pro-
gressed.  The prediction of the expansion of granular filters during
backwashihg usually was approached from an extension of filtration
hydraulics.

This extension of the fixed-bed hydraulics to a fluidized bed is
subject to challenge from a theoretical viewpoint.  In filtration or
fixed-bed conditions, the particles are not free to move about; while
in the fluidized state, they are suspended in the fluid and  free to
move about with little if any contact with other particles for two-
phase, liquid-solid fluidization.  However, expansion of the filter
media is seldom greater than 50%.  Because of this relatively low
degree of expansion, the extension of the fixed-bed hydraulics to the
fluidized bed has provided models that provide results which are
agreeable with the experimental results [24,29,46].

Recently, Amirtharajah [5] has shown  that the optimum porosity for
effective cleaning of silica sand by water backwash  is approximately
70%.  Expansion required to achieve this porosity would be dependent
upon the initial porosity ratio of the bed and the expansion-flow
rate characteristics of the graded media.  For a graded bed  of silica
sand, the required expansion would be approximately  45% to expand the
top  layer of  the bed, where most of the filtered solids are  retained,
to a porosity of about 707o.

Flow through  a Fixed Bed

Many of  the  sanitary engineering models for prediction of bed expan-
sion and  some of the expressions for  determining the minimum fluidi-
zation velocity of granular beds are  based in part on equations  de-
scribing flow through a fixed  granular bed.  Fair, Geyer,  and Okun
[46] present  the classical Kozeny  equation  for head  loss  through a
granular bed  for  laminar  flow  as,
            - u   (1 - e)2
             P8      3
                                  207

-------
where :

     h =  loss of head

     JL =  depth of bed

     k =  Kozeny's constant

     g =  acceleration due to  gravity

     jj, =  viscosity of the fluid

     p =s  mass density of the  fluid

     V =  superficial velocity of  the  fluid  above  the  bed

     e =  porosity

     d =  particle diameter  =  diameter of  equivalent volume  spheres

     \|l =  sphericity - defined as  the  ratio  of  the surface of  an
          equivalent volume  sphere to  the  actual surface area  of the
          particle.

From Coulson and Richardson [33 (p. 7)] ,  Carmen modified the  Kozeny
equation  to apply to transitional and turbulent flow.  This equation
is commonly called the Kozeny -Carmen  equation,
             2      •*     A    a
         p(V)     e3     d    8
where:
and
     R-     = drag force per unit area of particle  surface  in  the
              direction of flow motion


     V'     * — = velocity of the fluid in the pore openings
              £


     	i-j = 5 Re"1 + 0.4Re~0tl
     Re.    = modified Reynold's number = [Vp/(l - e)p,l [d/6] which
              uses specific surface (d/6) for the diameter  term.
                                208

-------
Another equation that describes head loss through fixed granular beds
in the laminar, transitional, and turbulent ranges was developed by
Ergun [23,40] (Clark, Viessman, and Hammer [29, p. 368] incorrectly
called Eq. (19) the Carmen-Kozeny equation).

                               2
                        3
                       ۥ
where:
     f- = dimensionless friction factor = 150 ^  " &' +1.75
      -L                                          Re
and
     Re = Reynold's number = ^—  .
                              P.

Sanitary Engineering Bed Expansion Models

In sanitary engineering practice, three different approaches used for
predicting expansion are of particular interest.

Two of the approaches are given by Fair, Geyer, and Okun  [46, Sect.
27, p. 19].

The Kozeny equation [Eq. (17)] and the constant head  loss equation
[Eq. (1)] can be equated and solved for the porosity  terms resulting
in,
        <

where:

     k  = Kozeny's  constant which  assumes  a value  at  about  4 for  a
          fluidized bed of low  expansion.

     Subscript  'i1  denotes the  'i'th  layer of  the  bed.

The  ratio of  the  expanded height,  ^ej>  to  tne  unexpended height,  j£c
is from Eq.  (2),

     *      A " ^
      C •
             l  -  e
                                  209

-------
 and the total expanded bed height Le is

                           

A log-log plot of Rec vs B produces a family of curves in which each
curve represents a different porosity.

The porosity of an expanded layer of the bed at a selected backwash-
ing rate is determined by trial and error as follows.  A trial poros
ity is selected, and Rec and B are calculated.  The intersection of
the Rec abscissa and the B ordinate on the family of curves is ob-
served to see if it falls on the curve representing that selected
porosity.  If not, a new porosity is selected and the process repeat
ed.  The expanded bed height is then calculated from Eq. (21).
                                 210

-------
Tesarik's discussion [130] of Camp's paper was critical of the group-
ing of terms for Camp's modified Reynold's number.  Tesarik redefined
the Reynold's number in a more conventional way,

          VPd
     Re a 	
           V>

where:

     Re = Reynold's number based on the superficial velocity (V) of
          the fluid above the bed.

Tesarik continued his discussion by showing that the expansion of a
granular filter is adequately described by Richardson and Zaki's
[100] expression of V/V^ = en (as presented below) and presented re-
sults of his own work.

        Bed Expansion Correlations from Fluidization Literature

Fluidization, although a relatively young field, gained most of its
importance in the early 1940's with the use of fluidized beds in the
catalytic cracking of petroleum.  Since that period, extensive stud-
ies of the fluidization process have been published.  The work of
Amirtharajah [4-6] is a collection of many of the different aspects
of fluidization that pertain to backwashing of granular filters in
sanitary engineering.

Richardson and Zaki's Correlations

Richardson and Zaki [100] determined that the ratio of superficial
velocity above the bed (V) to the settling velocity of a discrete
particle (Vs) is a function of the Reynold's  lumber (ReQ) based on
the settling velocity of  a discrete particle, porosity (e), and the
ratio of particle diameter to the tube diameter,
                    «. 51                                          (24)
where:
     — -  ratio  of  the  particle diameter  to  column  diameter.

Under  laminar and  turbulent  conditions,  the ratio  V/VS  is  indepen-
dent of
                                  211

-------
They presented their data graphically by plotting log V vs log e
(Fig. 56).  Above minimum fluidization, the data plotted as a straight
line with n representing the  slope of the line and V. the intercept
of the line at a porosity of  1.0, the mathematical expression for  the
line is
     log V = log V.  + n log  e
             (25)
or
They also observed empirically that Vs, for spherical particles,  and
V"i could be related by the expression
     log V
       0
^- 10
                                    d/D
             (26)
                   V. = VELOCITY INTERCEPT
                            n  SLOPE
        MINIA^JM FLUIDIZATION
        VELOCITY (V  .)  	
                    flflr
LOG
(SUPERFICIAL
VELOCITY, V)
                    LOG (POROSITY RATIO,
   Fig. 56.  Relationship between superficial velocity - porosity,
                               212

-------
combining Eqs. (25) and (26),


      V     en
     77- = —TTcr    for spherical particles only.                  (27)
      s   10Q/U

Replacing Vt for Vg in Eq.  (24) and from Eq.  (25), en  is  shown  to be

      n    V
     e  = —   for all particle shapes.                           (28)
          vi

Therefore, n slope is independent of  e and dependent upon the Rey-
nold's number of a free settling particle  (Re0)  and the ratio of  par-
ticle diameter to column diameter.  They developed the following  em-
pirical equations for n slope for spherical particles,

for

     0.2 < Re  < 1
             o

     n =  (4.35 + 17.5 |)Re~0-03                                   (29)


for

      1 < Re  < 200
           o

     n =  (4.45+18 j^Re"0'1                                     (30)


for

      200 < Re  < 500
              o

      n = 4.45 Re"0*1                                              (31)


The  above  equations  are  for the  transitional  range of flow.  Where
the  inertia  forces are negligible  (laminar regime),  the  results were
correlated by


      n =  4.65 +  19.5 |                                            (32)


and  where the viscous  forces are negligible (turbulent regime)

      n -  2.39.                                                    (33)

The  change in n  slope  for the three different flow regimes is  shown
graphically in Fig.  57.   In Richardson and Zaki's work,  they studied
                                  213

-------
           4.65
     n
    SLOPE
           2.39
                  LAMINAR
                  REGIME
                           TRANSITIONAL
                           REGIME
TURBULENT
REGIME
                            0.2
                                          500
          LOG
         REYNOLD'S  NUMBER,  Re  (BASED ON TERMINAL SETTLING
                       VELOCITY1, V,)

             Fig.  57.   Relationship between n slope and
                       Reynold's number, Re .
liquid-solid fluidization and sedimentation of spherical particles of
uniform size greater than 100 microns  in diameter with a density
range of 1.06 to 10.6,  in all three  flow regimes.

They had excellent correlations  between calculated Vs and observed Vj
as d/D -» 0, supporting  the validity  of Eq. (26), for spherical parti-
cles.

Wen and Yu's Correlations

Another expansion correlation was  proposed by Wen and Yu [138].  From
force considerations,  they showed  that porosity is a function of the
following equation,
f(e)
              F
               ks
                                                                (34)
where:

     F
      g
      gravitational  force
                             ,3
                           rrd pg
           buoyancy force =
                           TTd3pj
     F,   = drag force on a discrete particle
      KS

     g   = Newton's  law conversion factor.
                                214

-------
For the evaluation of F^g, they used Schiller and Naumann's equation,

            .2._2
           nd pv
                  (3 Re"1 + 0.45 Re"0'313)                        (35)
              ^>

which is valid for Reynold's number from 0.001 to 1000.

Using data from their own work as well as from the literature they
found, f(e) = e~4'7.  Combining Eqs. (34) and (35), the following
form of the equation was developed,


     e4'7 Ga » 18 Re + 2.70 Re1'687                              (36)

where:

          ^3  ,      N
          d p(ps - p)g
     Ga =       -      = Galileo number  (dimensionless).
Effect of Particle Shape on Expansion Correlations

The effect of particle shape on the expansion correlations has not
been studied extensively.

Richardson and Zaki's [100] work with nonspherical particles of regu-
lar shape (cylinder, cubes, plates) was in the turbulent range only.
They tried two different shape factors, sphericity (ty) and volumetric
shape factor (K).

The K factor correlated best for the evaluation of the n slope.  The
equation for n slope in the turbulent range was

     n = 2.7 K°'16                                                (37)

where:

            d3
     K  = ? —r = volumetric shape  factor
            d
             P

     d  - diameter of sphere of equivalent volume of  particle
      s

     d  = diameter of a circle of  same  area as the projected profile
      p   of a particle when lying in its most stable position.

Lewis and Bowerman  [79] reported that the expansion of nonspherical
particles could  be correlated by using  a modified porosity ratio  for
                                 215

-------
 the nonspherical particles.   The modification is obtained by multi-
 plying the  porosity ratio (e)  by the sphericity (ty)  of the particles
 OKe).

 Whitmore  [139,140], working with nonspherical particles,  found  that
 the n  slope of the  rough particles was  higher than Richardson and
 Zaki's n  slope for  spherical  particles  in the laminar range.  Whitmore
 stated that this higher n slope increased as  particle size decreased.
 He presumed that this  increase was because a  rough particle settling
 in a fluid  has a layer of attached fluid that smoothes off the  irregu-
 lar outline of the  particle.   This trapped liquid gives the particle
 a larger  effective  diameter and a lower density.

 Jottrand  [69]  fluidized uniform crushed sands with water  in the lami-
 nar range.   His plot of log V vs log e  resulted in a series of  paral-
 lel straight lines  with an n  slope of 5.60.   From the experimental in-
 vestigation, he found  that the fluidization velocity of given parti-
 cles at a given expanded porosity closely approximated the hindered
 settling  velocity of the fluidized bed  after  the  liquid flow was cut
 off.  In  subsequent experiments, he attempted to  replace  the  fluidi-
 zation experiments  by  a more  simple measure of the rate of hindered
 settling  of particles  after good agitation.   The  results  of the
 hindered  settling at a given  concentration of particles after good
 agitation was  found to be about 45% greater than  velocities observed
 by the first technique for the same particle  concentration.

 Wilhelm and Kwauk [141]  did extensive research on fluidization  with
 water  and air.   The particles  ranged in size  from 5  to 0.3 mm and in
 density from 1.125  to  10.792.   Their raw data of  flow rate, pressure
 drop,  and porosity  is  completely reported.  Of particular interest is
 their  work  with sea sand,  which was described as  prismatic with
 rounded edges.   Jottrand [69]  extended  the  analysis  of Wilhelm  and
 Kwauk's sea sand data  with a  log V vs log e plot.  The results  of the
 analysis  of the sea sands are  given in  Table  27.

 The  authors  of  the  present report calculated  the  n slopes for the sea
 sands  using Richardson and Zaki's equation  [Eq.  (30)],  and these
 values  are  included in Table 27.   The n slopes calculated from  Eq. (30)
 are  somewhat lower  than the n  slopes from the log V  vs  log e  plots
 for  the nonspherical particles as reported by Jottrand.

Carvalho  [28] ,  working with crushed anthracite coal  of  uniform  sizes,
plotted log V vs  log e  and determined the Vj[  intercept  and the  n
 slope.   He  also experimentally determined  the discrete  settling ve-
 locity, Vs,  for the various uniform sized coal particles  (Table 28).
The Vg  determined experimentally was  approximately 25%  lower  than the
V^ intercept (Table 28).   He attributed  this  discrepancy  to the shape
of the  crushed  particles.  Carvalho  did not compare  the n slopes from
the log V vs log e with  the n  slopes  calculated from Richardson and
Zaki's  equation [Eq. (30)].
                                  216

-------
Table 27.  Comparison of n slopes of sea sands (using Jottrand's
           analysis  [69] and Richardson and Zaki's equations).

Sand
1
2
3

Density,
g/cc
2.639
2.639
2.639

Diameter ,
mm Re,
0.373 21
0.556 45.5
1.000 122
Jottrand n
slope from
log V vs
log e plot
3.45
3.15
2.95

n slope
from
Eq. (30)b
3.28
3.04
2.76
 Reynold's number based on the velocity intercept at porosity ratio
  of  one  from  the log plot of V vs e.

  Assuming that Re  - ReQ and neglecting the d/D term.
Values of Rei from Carvalho's reported Vj[ and subsequent determina-
tions of n from Eq. (30) were calculated by the writers.  The results
are also presented in Table 28.

From the preceding literature, two significant points can be made:

1.   The n slopes for fluidized beds of nonspherical particles are
     greater than the n slopes of fluidized beds of equivalent sized
     spherical particles.

2.   The experimentally determined discrete settling velocity of a
     nonspherical particle of a bed does not equal the velocity in-
     tercept at porosity equal to unity of the log V vs log e plot.

Effect of Particle Size Distribution on Expansion Correlations

The media used for filtration does not consist of one-size  particles
but of a range of particle sizes.  The narrowest size range of media
that can be prepared conveniently and results in the largest to
smallest particle diameter ratio of about 1.2 is the media that is
retained between two adjacent sieves.  Although there is no limit to
the widest range of particle size, for multi-media filters the widest
diameter ratio suggested for any of the individual media comprising
the bed is about 2.5 [34].

Various investigators have used different methods of determining the
representative particle diameter for expansion correlations:
                                  217

-------
Table 28.  Comparison of n slope of crushed coal.






vi
fps
n
Average experimentally From log V
size, Temp, observed vs log e
Run3
I
II
III
ivd
V
VI
VII
vine
Runs
1.65
b., ,
mm
1.3
1.1
0.9
1.1
1.3
1.1
0.9
1.1
I to IV
g/cc.

°C
16
16
12
16
16
16
16
16
coal

_ j 	 j
.0
.0
.5
.0
.0
.0
.0
.0
0
0
0
0
0
0
0
0
density

l 	 L.

1_ _ __
fps
.194
.171
.139
.172
.208
.182
.165
.185
= 1.35


plot
0.
0.
0.
0.
0.
0.
0.
0.
g/cc,


265
255
176
203
278
232
222
220
Runs


4.20
4.70
4.70
4.15
3.70
4.00
4.35
3.80
V to



U
based on
V.
95
77
40
61
100
70
55
67
VIII


.2
.7
.3
.9
.0
.6
.5
.0
coal


n
from Re^ 'c
and
Eq. (30)
2.82
2.88
3.07
2.95
2.80
2.89
2.98
2.92
density =


°Assuming Re.  = Re  and neglecting the d/D term.

 Run IV the media for this run was comprised of equal volumes of media
 from Runs I to III.

^Run VIII the  media for this run was  comprised of equal volumes of
 media from Runs V to VII.
1.   In sanitary engineering practice  [24,29,46],  the particle diam-
     eter is described as the arithmetic mean of adjacent  sieve
     sizes.  Expansion correlations are made by summing  the  expansion
     of individual layers of filter media comprising a graded media
     bed.

2.   Amirtharajah [4,6], using data from the sieve analyses  of graded
     sands, plotted the percent passing by weight  on a probability
     scale against sieve size and used the diameter corresponding to
     60% passing by weight in his expansion correlations.  One sand
     investigated was not normally distributed about the mean size
                                  218

-------
     and did not plot as a straight line when plotted in this manner.
     This sand was analyzed as two separate components.

Wen and Yu [138] determined the effective diameter of two-component
mixtures by an inverse relationship,
     j
      inverse    n  W.
where:

     W. « weight fraction of  'i'th layer


     d. - mean diameter of weight fraction of  'i'th layer.

Van Heerden et al.  [134] , with gas fluidization of fine particles  in
the laminar range,  also used  the inverse definition of particle diam-
eter.  He compared  the d inverse, d  arithmetic  mean,  and d geometric
mean with minimum fluidization velocity studies and found that the d
inverse gave  the best correlation.   The arithmetic average gave
higher diameter values and the geometric average  gave lower values
than  the d inverse  diameter.

Leva  et al.  [77,78]  used the  arithmetic mean diameter (d_) ,

      d. = ZW.d                                                   (39)
      m      i i

for  two-,  three-,  and four-component mixtures  in  the  development  of
an equation  for minimum fluidization velocity.

Another method of  particle  size  determination  used by some  authors
[24,28,46] is the  equivalent  diameter of  a spherical  particle,
      *   =  =                                                    (40)
       eq
 where:

      Y  - particle specific weight
       s

      W  = total weight of N particles

      N  - number of particles.

 Fair, Geyer, and Okun [46] point out that when determining particle
 size by sieve analysis, the mean or 50% size is determined by weight,
                                  219

-------
 but the average diameter as determined by number is more closely rep-
 resented by the 10% finer size on a weight basis.

 Particle Segregation or Stratification

 Particle stratification resulting from backwashing is very apparent
 in rapid sand filters.  For equal density particles with varying di-
 ameters, different workers have reported rather widely different
 ratios above which stratification would occur.  These ratios have
 ranged from 1:1.3 [138]  up to 1:4 [96].

 Pruden [96,97]  did significant work to explain particle segregation
 in particulate  fluidization on a rational basis.  He presumed that
 the driving force towards segregation of two different groups of par-
 ticles is the difference in bulk density between the groups.  His
 definition of bulk density (pb) is,

      Pb = (1 -  e)pg + pe - (1 - e)(ps - p)  + p .                  (41)


 The bulk density difference between the large particles x and the
 small particles y is then,

                                                                  (42)


 Pruden used Richardson and  Zaki's  correlations for velocity  and  poros-
 ity [Eq.  (27)]  and  a theoretical equation for the settling velocity of,
 a  single  particle which  is  applicable in all ranges of flow,

           r> /     \ 1/m     1/m   ,(3-m)/n
           Cp(ps  - c\     - *    ' d^
where;

     C_. = drag coefficient  (function of Re )

     m  = index of fluid regime  (m = 1 in Stoke's regime, m = 2 in
          Newton's regime, m = 1.4 in transitional regime for a
          straight line approximation)

     n  = slope of the log V vs  log e plot.

Combining Eqs. (42) and (43), Pruden developed the following equation
for bulk density difference,
         * Pb
       x     y
Y
 b
                                       1 - r
                     nD
(3-m)/mn   10	
             (mn-l)/mn
           Yb
                     (44)
                                  220

-------
where:
          (% -p)
          d
           yr
     r  = T~ ratio of particle diameters.
           y

The assumptions of the above equation are that m^ - my = m, n^ = ny =
n, and drag coefficients are approximately equal.  He tested the
validity of the above equation experimentally by evaluating m, n, and
e from fluidization of the separate components.

He hypothesized that when the bulk density difference is equal to
zero, the components should be completely mixed.  A positive or neg-
ative value would mean that stratification would occur.  He experi-
mented with equal density and low diameter ratio components and con-
cluded that Eq. (44) does give the correct trend of component strati-
fication, but that a limiting bulk density difference for stratifica-
tion would depend upon properties of the particles.  He proposed the
following modification to the bulk density difference:  a reduced bulk
density, (3,
          (Pb
          V
                                                         (45)
           s
            x
where:

     0 £ p £ 0.01 mixing was observed

     0.01 < (3 £ 0.04 partial segregation occurred

     P > 0.04 segregation with  an  interface occurred.


Amirtharajah [4], in discussing Eq.  (44),  stated that  a  separate deter-
mination of ptx  and pfcy, by evaluating  the e  for each  component from
Eq.  (25) and solving for the bulk  density  difference in  Eq.  (42),
would eliminate  the uncertainty of some of the  assumptions.

Le Glair's thesis [74]  of two-component fluidization is  a  comple-
mentary study to Pruden's particle size segregation.   Le Glair ob-
served that when some  two-component mixtures  x  and  y,  such that
Do   > DO  and d  < d_,  were fluidized at low  flow rates, the x
rSx   roy      X    J
                                  221

-------
 component was below the y component.   At an increased flow rate, both
 components were intermixed.   At still a higher flow rate,  a reversal
 of the components occurred,  resulting in the small size,  dense mate-
 rial above the larger size,  less dense material.

 He explained that this phenomena could be attributed to the change
 of the individual bulk densities of the two components  as  flow rate
 changes.   At low fluidization rates,  the x component had  a greater
 bulk density than the y component,  and at increased flow  rates, the
 bulk densities were equal.   With still higher flow rates,  the x-
 component bulk density was  less than  the y-component bulk  density.
 Various densities and sizes  of media  can be selected which should
 display this behavior.   The  selection would depend upon the expansion-
 flow rate characteristics of the individual components.

 The porosity at which the particles are completely mixed  is called
 the inversion porosity (ei)  and is  approximated by,


                1 -  Yb
                                                          (46)
       I  "    1/mn     (3-m)/mn
          Yb     - ^    "    - Yb

which  is derived  from Eq.  (44) at  zero bulk density difference.

Le Clair pointed  out the following intermixing problems which are
of particular  interest.

1.  Equation (46) will give only an approximate value for ej and
    corresponding velocity because of all of the assumptions made
    for  the  development of Eq. (44) .

2.  Data collected for the individual components will not predict
    the  e-r precisely because  even  for a narrow size range, the in-
    dividual component data will give the average porosity of the
    component, not point porosities within the component.

3.  The  completely mixed state of  two components is reached at a
    lower velocity than predicted by the individual component data.
    This he  attributed to particle size distribution.  The small
    particles, sized by sieving, have a greater variation in size
    and  subsequently greater  range in bulk density than the larger
    particles, also sized by  sieving.  This difference in change of
    bulk densities would cause the completely mixed state to be
    achieved at a lower flow  rate than predicted from the individual
    component data.

4.  Inversion of the components would be observed for fluidization
    in the laminar range but not in the transitional or turbulent
                                 222

-------
     ranges because of turbulence, particle circulation, and macro-
     scopic mixing which would destroy the bulk density gradients.

 5.   The expanded height of a two-component mixture is closely ap-
     proximated by the sum of the expanded heights of the two in-
     dividual components whether the mixtures are segregated or
     ^/•mvnl Ai-01 IT ml -vo*1
          	A
completely mixed.
Another somewhat similar approach to intermixing is given by Camp
et al. [26].  They proposed that the driving force for intermixing
is the relationship between the particle density of the lighter ma-
terial and the bulk density of the more dense particles and water.

The equation used to calculate intermixing tendency was developed
by considering the forces acting upon the grains.  The buoyant force
(Fjj) on a grain in a f luidized bed is equal to the weight of the mix-
ture displaced,

     Fb = vpgpb                                           (47)

where:

     v  = the volume of mixture displaced

     p, = bulk mass density of the mixture.

The impelling force (Fj_) acting on a floating particle in the mix-
ture is its weight downward and the upward drag  force of  the wash-
water past  the particle.  The drag term presented  assumes spherical
shaped particles.
          weight
          term
where :
     CD  = drag coefficient  of the particle,  a function of Reynold's
          number and particle shape.

Equating F,  and F.  leads to the following:
                                 223

-------
 As  applied  to  a dual-media  filter,  if p^  of  the  lower heavier layer
 is  greater  than the  right-hand  terms for  the upper lighter layer,
 normal  separation of the  two media  will result.  If the reverse is
 true, intermixing can be  expected.  Camp  et  al.  presented very little
 experimental evidence in  support of the model presented above.

 However, using this  approach, they  concluded that at common back-
 washing flow rates of multi-media filters, the silica sand particles
 (1.00 to 0.595 mm) will fall into and mix with the garnet sand (0.500
 to  0.354 mm),  but that the  coal (1.41 to  1.00 mm) will be stratified
 above the silica sand.  The writers observe, however, that the coarse
 end of  the  coal is finer  than that  commonly  used in practice.

 Brosman and Malina,  in discussion of the  Camp paper [26], stated that
 intermixing for multi-media filters was more correctly described by
 the bulk density approach described by Le Clair  [74].  However, in
 their closing  discussion of the paper, Camp  et al. vigorously defended
 their model.

               Prediction of Settling Velocities

 The solution of the Richardson and Zaki expansion correlation, V/Vi e
 en, requires the velocity intercept of the log V - log e plot and the
 Reynold's number, corresponding to the settling  velocity (Vs) for the
 determination  of the  n slope.  Two different approaches can be used
 to find Vj.:  (1) the  direct determination from experimental observa-
 tion or from tables or graphs found in the literature, or (2) an in-
 direct method  of using the minimum fluidization  velocity and a ratio
 of settling to minimum fluidization velocities.

 Settling Velocities

Most textbooks in which there is a discussion of settling or sedi-
mentation give a method of calculating the settling velocity of a
 discrete spherical particle.  The settling velocity is solved by a
direct  solution in the laminar or turbulent  range.  In the transi-
 tional range, the calculation consists of a  trial and error solu-
 tion which  simultaneously satisfies the settling velocity equation,
 the drag equation, and empirical correlations of the drag coeffi-
 cient vs Reynold's number.  In the transitional  range, which is of
most interest, there  is considerable variation from the equations
 or graphs presented for the evaluation of drag coefficient of non-
 spherical particles.  Graphical methods of solving for settling
velocity of nonspherical particles suffer from the same weakness.

One particular method of solving for the settling velocity of a
discrete particle involves the Galileo number (Ga), a term which
occurs frequently in  the fluidization literature [17,33,53,69,76,
95,100,141],  The development for the Galileo number is as follows.
The resistance force  per unit projected area of  the particle when
equilibrium is established for a settling particle can be expressed as,
                                 224

-------
       Rl    J       J ^ X      V
        ^ rrd  = - rrd  (ps - p)g

or,
     R1 =   dg(ps - p)
where:

     R' = resistance force per unit projected area of the particle.
          Dividing both sides by pV2 and multiplying both by Re2
          yields,                  s                           °

      R'   2   2 d3
     ~2 Reo = 3 ~2  P(PS " P)g
     PVs         M.

             = | Ga.

The Galileo number is a dimensionless term that is independent of
velocity and the product of Reynold's number squared and drag force.
Interesting to note is that Ga is equivalent to Camp's dimensionless
backwashing number B, presented previously.

Coulson and Richardson [33] present a table and a graph relating Re
to Ga.  They also present a table for a correction factor to be
applied to the Re  for nonspherical particles.

Minimum Fluidization Velocities and Ratios of Settling and Minimum
Fluidization Velocities

Most of the formulas for predicting Vmf are developments from the
Kozeny equation [Eq. (17)] and the constant head loss equation
[Eq. (1)].  Leva [76] and Leva et al. [77,78] developed the following
by equating a modified Kozeny equation and the constant head loss
equation.  This equation incorporates a Reynold's number relationship
for emf and f,

           688 d1'82[Y(Ys - Y)]°-94

     Gmf	Of	                      (50)
                    M-

where:

     G f « superficial fluid mass velocity at minimum fluidiza-
           tion in Ib (mass)/hr sq ft

       d = diameter of particle in inches

    Y,Y  - fluid and particle specific weights in Ib/cu ft

       (j, = viscosity in centipoise
                                 225

-------
 valid for Remf < 10.  For Remf > 10, Leva presents a graphical cor-
 rection to be applied to G ~.

 Upon expressing Gmf as a superficial velocity (Vmf) ,  as done by
 Amirthara j ah [6] ,  the equation becomes,
                    ,1.82
      V   ' °-°0381
       mf    -                CK88
                            M"

 where:

      V ,. = minimum fluidization in gpm/sq ft

        d = particle diameter in mm

        p, = viscosity in centipoise,

 Again,  valid for Re f < 10.

 Zabrodsky [144] gives a multiplication correction  factor  for  the ve-
 locity  for Re  f >  10,


      k  - - 1.775 Re  "°*272                               (52)
      mf           mf                                    ^  '

 for  10 < Re , < 300.
            mf

 Wen  and  Yu [138],  starting with Ergun's equation [Eq.  (19)],  devel-
 oped  the following,
     Re
       mf
^(33.7)2 + 0.0408 Ga  -  33.7.                (53)
They used their own work plus the work of many others to develop
this equation which is valid for gas and liuqid fluidization with a
Re  ,. range of 0.001 to 4000.
  m£

Frantz [48], using gas-solid fluidization, made over 400 experiments
with eight different media and extensive analysis of the data deter-
mined that for the solution of the critical fluid mass velocity, (G f),
the theoretical coefficients and exponents give better results than
Leva's empirical correlations [Eq. (50)] and that further refinement
of the coefficients and exponents from the theoretical values was
not recommended for extrapolation outside the range of his experiments.
The theoretical equation for fluid mass velocity is,

                    5 d Y(YS - Y)
     G , = 4.45 x 10  	S	                        (54)
      mt                   j,
                                 226

-------
where:

       Gmf = fluid mass velocity in Ib (mass)/hr sq ft

         d = particle diameter in ft

      Y»YS = fluid and solid specific weights, respectively,
             in Ib/cu ft

         jj, = viscosity in Ib/hr ft


The relationship of free settling velocity and minimum fluidization
velocity has been presented by relating the ratio of Re /Rem£ to the
log Ga by several investigators [17,53,95],  Galileo number is a
constant for given particle and fluid properties.  Remf and Reo were
calculated from various empirical equations.  The results of the
various correlations of the ratio of Re^Re^ vs Ga are presented
graphically in the respective papers.

The above correlations could be used in the solution of the terminal
settling velocity by calculating Ga from the physical properties of
the fluid and the particle, and the Reo/Re^ ratio can be read from
the above correlations.  Remf can be determined from Eq. (50), (53),
or (54), or the appropriate equations in the following discussion.
Re0 can then be calculated from Remf and the Re0/Remf ratio.

A brief discussion of the individual papers follows.  Pinchbeck and
Popper [95] worked with gas-solid fluidization and plotted R
vs log 1/Ga.  Re0 was evaluated from,
     ReQ - - 6 + ^36 + | Ga.                            (55)

Van Heerden's equation for Re,^ [Eq. (56)] was chosen over Leva's
equation [Eq. (50)] because the former equation fit their data better,
     Re
-mf = 0.00123  (1 - emf) Ga.                         (56)
They assumed that emf was a constant of 0.406, as proposed by Van
Heerden et al.[134] for a bed of spheres of homogeneous diameter.
Hence, the constant (1 - emf) would be 0.594 in the above equation.
Pinchbeck and Popper's [951 correlations of Re_/Remf and Ga, using
limited experimental data, could be termed as fair.

Bourgeois and Grenier [17] used Ergun's equation [Eq.  (19)] and the
constant head loss equation [Eq. (1)] to develop the following equation

for Remf
                                  227

-------
      Re
        mf
                  (
1 + 3.11 x 10"4 Ga
                                                            O.5
                                                      *
                                                     - e  -)'
                                                       mf'
                                                          (57)
They  also  assumed em£  as  a constant 0.406, but pointed out the ef-
fects of a change in porosity  from this  assumed value and simplified
Eq. (57) further.  The  terminal  settling velocity was evaluated from
an empirical plot of Re  vs Ga.

They  also  found a substantial  difference of experimental results be-
tween air  and water fluidization and analyzed them separately.  They
developed  the following analytical expressions for the Re0/Remf vs Ga
correlation for water fluidization,
1.   50 < Ga < 2 x 10
     Re
       c
     Re
= 132.8 - 47.1 log Ga + 4.6 (log Ga)-
                             (58)
       mf
          Re
            mf
                <60
2.   2 X 104 < Ga < 106
     Re

     Re"
       mf
= 26.0 - 2.7 log Ga
                            (59)
         Re
            :
           mf
3.   10" < Ga
     Re
     Re
            9.0
                                               (60)
       mf
                                228

-------
Godard and Richardson [53] related Re f with Ga.  Starting with the
Kozeny-Carmen equation [Eq. (18)] andthe constant head loss equation
[Eq. (1)], they developed,
                     (1 - em£)          2    (1 - e-)0'1
     Ga - 180 Re _ • 	^-^- + 2.88 Re . • 	7^	      (61)
                mt       J              m£        J
                        Smf                      emf


They also used Ergun's equation [Eq. (19)] and the constant head loss
equation to develop,

                     (1 - e f)          7     ,
     Ga = 150 Re   • 	, m   +. 1.75 Re _ • -~               (62)
                mr       J              mr    J
                        emf                  emf


They related Ga to the Reynold's number based on the free falling
velocity by the following equations:
     Ga = 18 Re                      Ga < 3.6                  (63)


     Ga = 18 Re  + 2.7 Re1'687       3.6 < Ga <  105            (64)
               o         o


     Ga = \ Re2                      Ga > 105                  (65)
          J   o


Equations (63), (64), and (65) were obtained for the  laminar,  tran-
sitional, and turbulent range, respectively.  Equation  (64)  is ob-
tained from Schiller and Naumann's equation, previously discussed.

They also illustrated the effects of emf on the  Reo/Remf  ratio.  The
higher the emf, the lower the Reo/Re^ ratio for a given  Ga.  The
curves of Re0/Remf vs log Ga are quite sensitive to changes  of emf
at low values of Ga corresponding to the laminar range.   However,
this sensitivity diminishes as Ga increases and  is almost negligible
at high values of Ga > 10   (turbulent range).

The Re /Re   ratio reported in the preceding papers varied from 40
to 110 in Hie laminar range of flow, where the £„,£ effects on  the
ratio is most pronounced, down to 7 to 12 in the turbulent range of
flow.

Godard and Richardson [53] extended the use of the ratio  of  R
to the determination of the n slope.

Equation (27) can be rearranged to the following form of
                                 229

-------
     n = -.	 =	2_                               (66)
         log e    log  e f                                 v
 (if d/D  is negligible).  Using  this  relationship, it was possible to
 replace  Re Q/R^f with n  slope in  their correlations with Ga and to
 present  a series of curves of n slope vs Ga for different values of
 emf.  The values of the  n slope were somewhat higher than corresponding
 experimental values from the literature.  They attributed the high
 values of n slope to the phenomena that at minimum fluidization the
 particles become free to orientate in a manner to offer least resis-
 tance to flow, but the Ergun and  the Kozeny-Carmen equations do not
 reflect  this change and, therefore,  give higher values of Re^, which
 then give higher n slopes from  Eq. (66).

 Existing Models for Predicting the  Expansion of Fluidized Beds

 Amirtharajah* s Model

 Amirtharajah's [4,5] method for predicting bed expansion was a modifi-
 cation of the method proposed by Leva [76],  Amirtharajah's procedure
was,

 1.  Experimentally determine e  f  and p .
                              m£      s

 2.  From a probability plot of  the sieve analysis, determine the
    60% finer size (d.... -.    ).
                   v 60% finer7

 3.  Calculate V £ from Leva's equation [Eq. (51)]; if Re . > 10 apply
    Zabrodsky's correction factor, kmf, [Eq. (52)] to Vfflf.

4.  Use the relationship,
    to determine V .  The above expression is based on the empirical
    relationship of drag force on an isolated spherical particle to
    drag force on the same spherical particle in a fixed bed,
        £
          particle isolated _ 71 3                       /67\

          particle in bed

    This ratio was proposed by Rowe [105]  and Rowe and Kenwood [106]
    and validated by Davies and Richardson [36].  Amirtharajah also
    assumed that drag forces are proportional to the square of the
    velocities; thus he used the square root of the drag force ratio
    in the above expression.
                                  230

-------
5.  The n slope can then be determined by the use of Richardson
    and Zaki's equations [Eqs. (29) through (33)] where Reynold's
    number is based on V .
                        s
6.  Using a modified form of Eq. (27),

            V = ke11                                      (68)

    and inserting V^ for V, efflf for e, and n into the equation,
    the value k can oe determined.  The k value has the same units
    as a velocity term.

7.  The e of the expanded bed can then be determined for any super-
    ficial velocity, V, by reapplying Eq. (68) with the values of k
    and n slope previously calculated.

8.  The expanded bed height is then found from Eq. (2).

Step 4 of this model is incorrect.  The ratio of drag forces used by
Amirtharajah (71.3) is valid for rhombohedral packing of spheres and
can thus be reasonably applied to the bed at the onset of fluidization.
However, from fundamental hydrodynamics, the drag force is proportional
to V  in laminar range, V2 in turbulent range, and between vl and V2
over the transitional range.  Thus, Amirtharajah's use of drag force
being proportional to V2 is the source of his error and is a misinter-
pretation of Rowe's paper [105].  This error was somewhat self-corrected
in Amirtharajah's subsequent steps.  For Amirtharajah1s fine sand A,
his reported value of Ga was equal to 5810 [4, p. 109].  The ratio of
ReQ/Re f or VS/Vf for this Ga should be approximately 20 from the
literature [19,53,95] rather than V 71.3 as used by Amirtharajah.  This
would mean that the Vs and the Reo Amirtharajah used for determination
of the n slope from Eq. (30), neglecting d/D, would be lower by a fac-
tor of 20/Y 71.3 <=- 2.5.  The correct n slope would, therefore, be larger
by a factor of approximately,


     nAmirtharaiah _    4.45(ReQ)        ^       ^
                   -                 n i ~ ^  • '    -i.uy.
      "correct       4.45(ReQ x 2.5) "*


In his step 6, e f, V  -, and n slope were used to calculate a k value.
Theoretically, for spnerical particles, this k value should approxi-
mate^.  In his thesis [4, p. 113], the Vg calculated from Vs/Vmf =
Y71.3 and the k value  from subsequent calculations were 75.1 gpm/sq ft
and 140.7 gpm/sq ft, respectively.  The ratio of these values is
140.7/75.1 =- 2, which  is approximately the factor which the Vg/V^ or
Re /Rem£ ratio was assumed to be off.  Thus, Amirtharajah used  an n
slope which was slightly high, which led to an erroneous k value.  Thus,
his expansion model was somewhat  self-correcting and gave good  results
when compared with his experimental data.
                                  231

-------
 Leva's Model

 Leva's method of predicting bed expansion is as follows:

 1.  Experimentally determine e f and pg.

 2.  From the sieve analysis, define the diameter by the inverse
     definition [Eq. (38)].

 3.  Determine the minimum fluidization mass flow rate,  G ,.,  from
     Eq. (50).                                           raf

 4.  Determine the fluid mass flow rate, Gf, at the superficial
     fluidization velocity of interest,

 5.  From this G^, determine the Reynold's number using  Vf  correspond-
     ing to Gf, and from Richardson and Zaki's equations (Eqs.  (29)
     through (33)] determine n slope.

 6.  Calculate V.  from the previously determined Vm£,  e_f,  n  slope,
     and Richardson and Zaki's Eq.  (28), V,  = V J(e £)n.
                                          l    mi   mr

 7.  Repeat the above  calculation using V.,  n slope,  and V  to solve
     for the expanded  e at Gf.

 8.  The expanded  height can be determined by Eq. (2).

 Amirtharajah [4,5]  pointed out a significant error in Leva's model.
 Leva uses  the Reynold's number based  on the superficial flow rate of
 the  expanded bed  to determine the  n slope.   This is  incorrect.
 Richardson and Zaki's  n slopes should be determined  from the Reynold's
 number  based on Vi  at  e of 1.0.  For  spherical particles,  this would
 be the  point where  V   = V..

 Wen  and Yu's Method

 Wen  and Yu's [138] method of  predicting the  expansion  is very straight-
 forward.  The expanded porosity  can be determined  by  the fluid and
 particle properties and Eq.  (36),


      e4'7Ga = 18  Re +  2.70 Re1'687  .

 The  expanded height of the bed can  then be  calculated from Eq. (2),
     * ~  (1 - e)

S,  and e  having been previously determined.
                                  232

-------
       X.  EXPANSION AND INTERMIXING EXPERIMENTAL INVESTIGATION

                       Experimental Apparatus

General Layout of 6-in. Fluidization Column

A schematic layout of the 6-in. fluidization apparatus is shown in
Fig. 58.  This apparatus was used in the expansion vs flow rate ob-
servations of garnet sand, silica sand, and anthracite coal for
purposes of developing expansion prediction models.  It was also
used in observations of intermixing of silica sand and coal and the
hydraulic gradients which exist in these intermixed beds.

The source of water (hot and cold) was the university tap water
supply.  Two sets of water taps were used, A and B.  The high flow
rates were metered by flowmeter F]_.  The lower flow rates were passed
through a thermostatically controlled mixing value, C, then through
flowmeters Fo and F3 and to the fluidization column, D.  The water
temperature was measured by a thermometer placed within the expansion
column.

The fluidizing column consisted of 6-in. inner diameter, 1/2-in. thick
plexiglass tube 4 ft 5 in. deep with a 3-in. high calming section at
the bottom.  The water was fed through 59 orifices of 1/16-in. diameter
in a 1-in. thick underdrain plexiglass plate.  Sets of orifices were
staggered from one another so as to provide a uniform matrix of ori-
fices on the entire plate.  The calming section was filled with 1/2-in.
diameter glass marbles.

The solid particles composing the bed were supported on two stainless
steel meshes  (No. 50 over No. 10) placed above the 1-in. plexiglass
plate with the orifices.  Pressure taps were located on the column
to permit observation  of pressure drop along the depth of the bed.

The first pressure tap was placed  in the bottom  flange of the column
and projected to within 1/4 in. above  the  stainless steel screens.
The second pressure tap was placed in  the  column wall 3 in. above  the
screens and directly above the first pressure tap.  The rest of the
pressure taps were at  3-in. intervals  up  the column in the garnet
sand experiments.  An  identical column with pressure  taps  at 1.5-in.
intervals was used on  the silica  sand  and  coal experiments.  The
pressure taps were constructed of  1/4-in.  copper tubes.  The inner
opening of the pressure  taps was  covered by  a 50-mesh or 100-mesh
 stainless  steel  screen soldered  in place.  The first  pressure  tap
protruded  1-1/2  in.  into  the column.   The remainder  of  the pressure
 taps protruded 1/2  in. into  the  column.

The above  pressure  taps  were connected by plastic tubing to  glass
piezometers mounted  on boards.
                                   233

-------
CO
           UNIVERSITY WATER SUPPLY
                 A           B
             HOT  COLD  HOT  COLD
            FLUIDIZATION
            COLUMN D
                                                                              PIEZOMETER BOARDS
         FLOW
         METERS
                                THERMOS! AT 1C
                                CONTROL VALVE
                                C
PRESSURE
TAPS
                                    -*	M-
                                              E'
                          Fig. 58.  Schematic layout  of 6-in. fluidization column.

-------
General Layout of 2-in. Fluidization Column

A schematic layout of the intermixing column is shown in Fig.  59.
This column was used in expansion vs flow rate observations for
uniform sizes of silica sand, garnet sand, and coal.   It was also
used in observation of intermixing tendency between dual media com-
prised of uniform sizes of either garnet and silica sand or silica
sand and coal.

The source of water (hot and cold) was the water taps, B, discussed
in connection with Fig. 58.  The mixing valve and flowmeter F, were
also used with the 6-in. column experiments.  Flowmeter F^ was
connected in series to flowmeter Fo.  Between the two flowmeters was
a 1/4-in. needle valve and a 1/4-in. quick shutoff valve.

The column consisted of the calming section, the fluidizing chamber,
and an overflow structure.  The fluidizing chamber was made of a 2-in.
inside diameter plexiglass column 68-3/4 in. in height with 3/4-in.
plexiglass flanges on each end.  The inlet to the calming section was
by a 3/8-in. opening.  The height of the calming section was 2-1/4 in.,
with an inside diameter of 2 in.  The lower 1-1/2 in. were filled with
glass beads 6 mm in diameter.  A stainless steel screen of 50 mesh was
placed in the bottom of this section to prevent loss of the glass beads,
The remaining volume of the calming section was filled with 2-mm lead
beads.  Between the calming section and the fluidizing chamber, a
100-mesh stainless steel screen was placed.

The overflow structure was an 11-1/4-in. extension of the fluidizing
chamber.  The top of this structure was perpendicular to the column
axes, and the water flowed radially out of the column.  This water
was collected by a circular trough and drained to the floor drain.

Flowmeters

Four flowmeters were used during the collection of data.  They were
the rotameter type and, for the purpose of this research, were desig-
nated as F,, F-, F», and F, .  The range of flows and  scale of the
flowmeters  are as follows:

Flowmeter


    Fl

    F2

    F3

    F,
                                  235

-------
UNIVERSITY WATER SUPPLY
          B
      HOT  COLD
             u
             I i
OVERFLOW
STRUCTURE
                         FLUIDIZATION
          L                 CO
         (yjTHERMOSTATIC
         ^-^ONTROL VALVE
             FLOW METERS
                      RUBBER HOSE!








5N
rtN
g




N


1
S<
                                          '  SCALE
                                            LEAD BEADS
                                            GLASS BEADS
  Fig. 59.  Schematic layout of 2-in. fludization column.
                          236

-------
Flowmeters F]_, F2, and F3 were calibrated at the two temperatures
used by a weighting technique.  A least squares fit was used in
determining the calibration equation for the flowmeters.  These
equations are as follows:

Flowmeter    Temperature.  C             Actual flow,  gpm

   FX              16.5         • 0.038 + 1.007 (meter reading) X 24

                   25.0         =0.113+1.003 (meter reading)  X 24

   F2              17.0         =0.169+1.032 (meter reading)

                   25.0         =0.197+1.027 (meter reading)

   F3              17.0         = - 0.067 + 0.994 (meter reading)

                   25.0         - - 0.069 + 1.026 (meter reading)

The flowmeter reading of meter F, was checked using a volumetric
technique.  The results of this calibration were judged acceptably
close so that the flowmeter reading was accepted to be the actual
flowrate.

Sieves

United States Standard sieves were used in determining the gradation
of the sands and also to separate uniform sizes from wider size ranges
received from the suppliers of the garnet sand, silica sand, and coal.

Filter Media

Two different types of sands, garnet and silica, and crushed anthracite
coal were used in this research.  The garnet sands are obtained from
alluvial deposits which are crushed to requirements and sieved to de-
sired specifications.  The individual particles are very angular and
vary widely in shape.  The recommended sizes of garnet sand for waste-
water filtration range from passing United States Standard Sieve No.
20 to that retained on No. 80 134],

Five different garnet media  (Idaho Garnet Abrasive Company, Kellogg,
Idaho) were studied in the garnet expansion experiments.  They were:
(1) uniform sized (-14+16) (uniform sizes are defined  as passing (-)
and retained United States Standard sieves) separated  from manufactur-
ers rating size M-16, (2) uniform sized -25+30  separated  from manu-
facturers rating  size M-36,  (3) uniform sized -50+60 separated from
manufacturers rating size M-60-80, (4) graded sized, as received,
manufacturers size M-60-80 and,  (5) graded size, as received, manu-
facturers rating  size M-36.
                                  237

-------
 For the garnet-silica sand intermixing experiments, a single uniform
 garnet sand was fluidized in successive experiments with various
 uniform silica sands.  The garnet sand was the uniform -50+60 used
 in the expansion experiments.  The silica sand used in this series
 of experiments were uniform sizes of -20+25, -30+35, -35+40, and
 -40+45 separated from manufacturers (Northern Gravel Co., Muscatine,
 Iowa) rating fine sand.  The graded silica sands and anthracite coals
 used in the 6-in. column experiments are presented in Table 29.


 Table 29.   Size and source of raw graded silica sands  and coals
            studied.
Type
Graded sand
Graded sand
Graded coal
Graded coal
Graded coal
Graded coal
Graded coal
Graded coal
Designation
A
C
A
B
C
D
E
F
Effective
size,
mm
0.54
0.72
0.89
0.98
1.08
1.30
1.46
1.92
Uniformity
coefficients
1.41
1.42
1.53
1.43
1.53
1.47
1.38
1.72
Source
1
1
2
2
3
4
4
3
 1.   Silica  Sand, Northern Gravel Co., Muscatine,  Iowa.

 2.   Philterkol No.  1  Special Anthracite, Reading  Anthracite Coal
     Company,  Locust Summit, Pennsylvania.

 3.   Carbonite #B, Carbonite Filter Corp., Delano, Pennsylvania.

 4.   Shamokin  Coal Company, Shamokin, Pennsylvania.

Uniform silica sands  and coals were used in the 2-in. column to ob-
 serve expansion, bulk density, and intermixing behavior as a function
of flow rate.  These uniform media were prepared by sieving the above
graded media  and included the following sizes:  -10+12 and -12+14 mesh
 from Sand C;  -14+16, -16+18, -18+20, -20+25, -30+35, -35+40, and
 -40+45 mesh from Sand A; -4+7, -7+8, -8+10 and -10+12 mesh from Coal
F; and -10+12, -12+14, -14+16, '16+18, -18+20, -20+25, and -25+30
mesh from Coal A.
                                  238

-------
                       Experimental Procedures

Separation Sieving for Uniform Media

The two sands and the coal that were used were sieved into uniform
sizes by using a stack of appropriate sieves and a Combs Gyratory
machine (Great Western Mfg. Co., Combs Gyratory Sifting Machine,
Leavenworth, Kansas).  The sieving time was 5 min for sand and 8
min for coal, and the total load to the sieves was 300 to 350 g
for sand and 200 to 250 g for coal.  The sands and coals of each
uniform size were sieved a second time using the same procedure as
in the first separation sieving to improve the uniformity.

Average Particle Size Determination

Sieve analysis.  The ASTM Standard Method for Sieve or Screen
Analysis of Fine and Coarse Aggregates [2, p. 93] states the fol-
lowing criteria for sieve analysis:

     The sample of aggregate to be tested for sieve analysis
     shall be thoroughly mixed and reduced by use of a sample
     splitter or by quartering — to an amount suitable for
     testing.

     ... not more than 1% by weight of the residue on any
     individual sieve will pass that sieve performed as follows:
     Hold the individual sieve, provided with a snug-fitting
     pan and cover, in a slightly inclined position in one hand.
     Strike the side of the sieve sharply and with an upward
     motion against the heel of the other hand at a rate of about
     150 times/min, turn the sieve about 1/6 of a revolution at
     intervals of about 25 strokes.

The criteria of not more than 1% passing in an additional minute
of hand sieving was not met by using the sieving machine.  Because
of this, machine sieving was abandoned for the sieve analysis pro-
cedure, and a modified method of hand sieving was adopted.  This
method followed the above ASTM Standard Methods describing hand
sieving except that the sieving time was extended to 3 min.  It is
recognized that the hand sieving technique would also have been
better for the separation  sieving previously described.  However,
hand sieving was not practical for that step due to the large
quantities of media  that were separated.

Equivalent diameter of a sphere.  The particle size of  the three
uniform garnets was  also determined by counting  and then weighing
particles and calculating  the equivalent diameter of a  spherical
particle of the  same weight and density.  The equivalent diameter
of a spherical particle is given by Eq.  (40),
                                239

-------
      d
       eq.


 Density

 The densities of the silica and garnet sands were determined by the
 water displacement technique using a 50-ml pycnometer bottle.  The
 sand samples were dried at 110 °C for 2-1/2 hr as a preliminary pro-
 cedure prior to the test and were submerged for 1 hr before final
 weighings.

 The same water displacement method was used for the coal because the
 filter coal media during actual operating conditions would be
 similarly submerged in water.  The samples were dried at 110 °C for
 2-1/2 hr as a preliminary procedure prior to the test and were  sub-
 merged for  24 hr before final weighings to allow the water to pene-
 trate the pore spaces.

 Porosity

 Two methods of porosity determination were used.  The first was a
 water displacement and  simulated fluidization technique, hereafter
 referred to as the graduate cylinder technique.   Two 1000-ml graduate
 cylinders were used.  In one cylinder, sand with a dry volume between
 200 and  400 ml was measured.   In the other cylinder,  exactly 500 ml
 of  water was placed.  The known volume of sand was poured slowly into
 the cylinder that contained the 500 ml of water.  The total volume of
 the sand and water and  the apparent volume of the sand were measured.
 The next step was to  simulate fluidization of the particles.  But
 before this was done, enough water was added to completely fill the
 cylinder.   There were two reasons  for adding the additional water:
 (1)  to prevent the trapping of air in the settled bed and (2) to pre-
 vent some of the particles from sticking  to the  sides of the cylinder
 and rubber  stopper.  With a rubber stopper placed tightly in the open
 end of the  cylinder, the cylinder  was rapidly inverted a number of times
 then quickly set down and the particles allowed  to settle.   The apparent
 volume of sand was then measured.   In this  method,  it was assumed that
 the particles would settle in their least dense  volume and represent
 the same porosity as that at  minimum fluidization velocity.  Three
 sets of measurements were made for each garnet-sand media.

The  second method used  for porosity determination of  a media in a
 fixed-bed condition in  the column  will be hereafter referred  to as
 the  column  technique.   The volume  of media  was found  from the weight
 of  the media placed in  the fluidizing column and the  specific weight
 of  the media.   The total  volume  of the media and entrapped fluid  was
determined  from the column diameter  and the bed  height of the media
 after the bed was  fluidized,  expanded, and  slowly contracted.   The
column technique was used  as  a check of the graduate  cylinder technique
 for  the garnet  sand and was the  only  method used for  porosity deter-
minations of the  silica  sand  and coal.
                                  240

-------
Settling Velocities

The settling velocities of discrete particles were experimentally
determined for representative samples of the three uniform size
garnet sands.  A 5-in. diameter plexiglass column, 56 in. in height,
was used for this experiment.  The column was filled to the top with
water.  The particles that were placed in the water, rolled gently
between two fingers to completely remove any air attached to the
particle, and then released.  They then fell through 16 in. of water
to come to dynamic equilibrium before a stopwatch was started.  The
particles then fell through a 30-in. timed interval.  The average
settling velocity of each uniform garnet-sand media was determined
at two different water temperatures of 16 to 17 °C and 25 °C.  The
water temperature was adjusted when it deviated more than 0.5 °C from
the temperature desired.

Expansion — Flow Rate Experiments

Expansion experiments were made in both the 6-in. column and the
2-in. column previously described.  The total bed was fluidized to
about 50% expansion and allowed to contract slowly.  Starting with
a fixed bed, the flow rate was incrementally increased to a maximum
expansion.  Readings were taken of flow rate, temperature, bed height,
and (for the 6-in. column) the piezometer tubes.

Readings were not made until the water temperature was constant, the
influent temperature matched the effluent temperature, the water
pressure was steady, and the bed height was stabilized.  Visual
observations of bed behavior, such as portions of the bed which were
fluidized, and any circulation patterns were also noted.

Upon fluidization of the first garnet sample, run 1 (Series A-13), it
was noticed that the top 1/2 to 1 in. of the bed consisted of a lighter
colored material than the rest of the bed, which was the purple
color characteristic of the garnet sand.  The bed was completely
fluidized and then contracted slowly.  The light-colored greyish
material was then siphoned off the top of the bed.  Removal of this
light-colored, less-dense material was done before any expansion data
were taken for garnet gradations.

Bulk Density and Intermixing of Uniform Sized Media

A full range of uniform-sized media of each type (coal, silica sand,
and garnet sand) were fluidized in the 2-in. column for purposes of
determining expansion vs flow rate.  From these data, bulk density
could be calculated at all flow rates.  Some of the uniform media
were then observed in dual-media beds comprised of silica sand and
coal or garnet sand to determine their intermixing behavior.  The
procedure for the garnet sand-silica sand experiments is presented
here as an example.  Essentially the same procedures were used for
the coal and silica sand observations.
                                  241

-------
 One uniform size garnet sand -50+60 and uniform size ranges of silica
 sand -20+25, -30+35, -35+40, and -45+50 were used.  The garnet was
 split down to approximately 1000 ± 40 g samples by the use of a sample
 splitter.  The samples were then adjusted by the removal or addition
 of garnet to 1000.0 g ± 0.1.

 The silica sand was separated into uniform sizes as described pre-
 viously, washed, and dried.  Then the sieve analyses were made.   The
 uniform silica sand samples were then weighed with a precision of
 ± 0.1 g.

 One of the 1000-g garnet sand samples and all of the uniform silica
 sand samples were individually fluidized in the intermixing column.
 The samples were completely fluidized and slowly contracted to a
 fixed bed at zero upward flow.  Then the flow rate was incrementally
 increased.  Readings of temperature,  bed height, and flow rate were
 recorded during the expansion.  Visual observations of the portions
 of the bed fluidized and the circulation patterns were also recorded.

 After each of the individual silica sand media was fluidized and
 observed as described in the above paragraph,  a 1000.0-g garnet  sand
 sample was poured into the  intermixing column on top of the silica
 sand.   The dual media was then expanded 60  to  70% and  very slowly
 allowed to contract to a fixed bed with no  flow.   Bed  height,  relative
 location of each media,  and qualitative concentration  of each  media
 were noted.   Expansion of the  bed  was started  and incrementally  in-
 creased to a height of about 60 in.  (200%).  The previously mentioned
 bed height,  qualitative location and  concentration of  the individual
 media,  temperature,  and visual observations  of circulation patterns
 were recorded at several flow  rates.

                       Illustrative Calculations

 The  illustrative calculations  presented are  for  the  uniform garnet
 sand -14+16 which had  been  separated  in accordance with the separation
 sieving technique previously described.

 Sieve Analysis

The  results of  three  sieve  analyses are shown  in Table  30.  The  results
of using  the mechanical  sieving machine and hand  sieving  are shown.
 Sample  1 was  sieved  for  5 min  on the  Gyratory  sieving machine.   Samples
2 and 3 were hand sieved  for 3 min.   Samples la  and  3a  are  the per-
centage of the material  retained on an  individual sieve which passes
that sieve in one additional minute of  hand  sieving.  The percentage
 that passed the No. 14  sieve in the additional minute of hand  sieving
is highly distorted because  the amount  that was  retained  in the  initial
sieving was very  small.  The percentage  that passed  the No.  16 sieve
in the  additional minute of hand sieving  is slightly higher than  the
ASTM recommended value of 1% passing  an  individual sieve  in one  addi-
tional minute of hand  sieving, but comparison of  the mechanical  and
                                  242

-------
Table 30. Sieve analysis of garnet sand media (-14+16).
Sample
no.
Sieved
load, g
Sieving
procedure
Sieving
time, min
Sieve
no.
14
16
18
Pan
1
148.99
Machine
5
Sieve 7.
opening, retained
mm
1.41 16.72
1.19 81.32
1.00 1.85
0.11
la
Hand
1
7»
passing3
23.20
11.54
0.08
—
2 3 3a
145.92 154.15
Hand Hand Hand
331
7o 7. 7o
retained retained passing3
0.17 0.31 16.66
88.10 87.35 2.38
11.69 12.30 0.00
0.04 0.04



7, retained
mean of
2 and 3
0.24
87.72
12.00
0.04



% passing
mean of
2 and 3b
99.76
12.04
0.04
0.00
 Percent passing in one additional minute of hand  sieving  expressed as percentage of original amount
 retained on that individual sieve.

30ne hundred minus mean percent retained.

-------
 hand sieving methods  shows the improvement of the hand  sieving method
 over the mechanical sieving method.   The results of sample  1 are  rep-
 resentative of the results obtained  by mechanical sieve analysis  of
 other samples for sieving times of up to ten minutes.   The  increase
 in mechanical sieving time was found to improve  the sieve analysis
 only slightly.   Because  the results  of hand sieving method  conform
 quite closely to the  recommended standards of ASTM,  as  previously
 stated,  the hand sieving method was  adopted,  and all sieve  analyses
 reported herein are by the 3-min hand sieving technique.

 Average  Particle Size Determination

 From sieve  analysis.   The average sizes  are determined  by the  inverse
 definition  Eq.  (38) and  by the arithmetic  mean definition Eq.  (39)
 as follows:

                         Mean opening    Weight
 Sieve Sieve opening,  between sieves   fraction
no.
12

14

16

18

Pan


mm
1.68

1.41

1.19

1.00

0.841


(di), mm

1.55

1.30

1.09

0.92



(wt)

0.24

87.72

12.00

0.04

W.
~ JL
di
Wi/di

0.16

67.48

11.01

0.04


78.69
Widi

0.37

114.04

13.08

0.04



                                                  I Wtd  - 127.53


Therefore, changing of the weight from percent to a fraction, the
diameter as defined by the inverse definition Eq. (38) is,
     Inverse = "iT = 0^869 = U271
and as defined by the arithmetic mean average from Eq. (39) is,


     d_ « SW.d   =  1.275 mm
      m
                                  244

-------
Note:  the difference in the two definitions of diameter is small for
a uniform media but is greater as the variation in size of media in-
creases as will be shown in the results section of this report.

Equivalent diameter of a sphere.

        Sample                           I                     2

Number of particles                  110                   110

Weight of particles                    0.5798                0.5968

Density of solid, g/cc                 4.140                 4.140

Equivalent diameter of a spherical particle of the same volume is given
by Eq. (40),
d
 eq
                    [6(0.5798") "I1/3                [6(0.5968)
                    |_110 TT 4.15J                   [110 TT 4.15J
     de (mm) =          1.343                          1.356



     avg d  (mm) -                       1.349


Density

 1. Weight of dry pycnometer                             =  25.9310 g

 2. Weight of pycnometer full of water                   = 125.4432 g

 3. Weight of pycnometer with inside wet                 «  26.1743 g

 4. Weight of pycnometer inside wet + garnet sand        «  45.7310 g

 5. Weight of pycnometer + garnet sand + water to fill   = 140.2815 g

 6. Temperature of water                                 =    24  C

 7. Therefore, weight of sand =4-3                    =  19.5567 g

 8. Weight of water to fill pycnometer - 2 - 1           =  99.5122 g

 9. Weight of extra water to fill pycnometer over the    =  94.7938 g
    sand =5-1-7

10. Weight of equivalent volume of water - 8 - 9         -   4.7184 g
                                   245

-------
 11. Therefore,  specific gravity of garnet  sand  at 24 °C  = 7/10
     =  4.1448

 12. Density of water at 24 °C                            = 0.99707 g/ml

 13. Density of garnet sand = (11) (12)                   «, 4.133 g/ml


 Porosity

 Graduate cylinder technique.  The following three samples are from the
 garnet sand -14+16,

                                           i          ii       iii

 1.  Dry volume of sand (ml)             340       390       170

 2.  Volume of water (ml)                500       500       500

 3.  Total volume of sand + water (ml)   695       740       610

 4.  Volume of sand after simulating     360       440       200
     fluidization (ml)

 5.  Porosity [4 - (3-2)]/4                0.458      0.455     0.450

     Average of row 5 =  0.454



 Column  technique.

 Dry weight of  -14+16 garnet  sand removed from  column,  Ib  *   31.5

 Fixed bed  height,  ft                                     =    1.138

 Cross-sectional  area of  column, sq  ft                    »    0.196

 Total volume of water and  garnet sand,  cu ft              =    0.2235

 Specific weight of garnet  sand, Ib/cu ft                 = 257.9

 Volume of  garnet  sand removed from column,  cu  ft         =    0.1221

 Fixed bed porosity - (0.2235 - 0.1221)/O.2235             -    0.453

 Settling Velocities

Number of particles dropped                               * 100

Distance of timed fall, ft                                =2.5

Temperature of water                                      « 16.5  C
                                  246

-------
           Observed settling time of single particles, sec
3.3
4.9
3.4
2.8
3.8
4.4
3.4
3.9
3.6
3.4
Mean
3.
2.
3.
3.
2.
4.
3.
3.
4.
3.
time
Standard
Mean
5
9
6
0
7
3
7
1
0
9
of
3.5
3.2
4.4
2.9
2.6
4.0
3.1
3.1
3.2
3.7
fall
4.1
3.2
3.0
3.0
3.1
3.5
4.4
3.4
3.0
3.6

3.1
3.0
2.9
3.0
4.2
4.3
4.3
3.7
3.5
3.8

2.9
3.0
4.1
3.4
3.5
3.7
3.6
3.0
4.0
4.4

3.9
3.2
3.3
3.0
3.1
3.1
3.5
3.2
3.8
4.3
=
deviation =
velocity =
0.703
fps =
3
.4
3.1
5
3
4
2
3
3
4
4
3.56
.4
.9
.0
.9
.2
.7
.2
.0
sec
3.4
4.2
3.1
3.1
3.2
3.4
4.1
3.8
5.1
3.9

4.2
3.6
3.2
3.2
4.8
3.1
2.8
3.7
3.5
3.2

0.5534 sec
315
gpm/sq
ft

Expansion — Flow Rate and Intermixing

Illustrative calculations for the expansion-flow rate and intermixing
data will be presented along with the presentation and analysis of
the data.
                 Results and Analysis — Summary

The main series of experimental runs for the study have been summarized
in Tables 31-32.  Table 31 lists the downflow and/or upflow runs in the
large column for various single media.  Table 32 describes the upflow
and downflow runs in the large column for various dual-media filters.
Table 33 gives the upflow runs in the small column for uniform single-
and dual-media filters.

                   Results — Media Characteristics

Sieve Analyses

The results of the sieve analyses for the various media studied are
presented in Figs. 60, 61 and 62.  Effective size and uniformity
                                  247

-------
Table 31.  Upflow and/or downflow experimental runs with the various single
           media in the 6-in. column.
Series Media3
A-l
A-2
A-3
A-4
A- 5
A-6
A-7
A-8
A- 9
A-10
A-ll
A-12
A-13
A-14
A-l 5
A-16
A-17
Coal A
Coal UCA
Coal B
Coal C
Coal C2
Coal D
Coal E
Coal F
Sand A
Sand A£
Sand C
Sand C2
Garnet
-14+16
Garnet
-25+30
Garnet
-50+60
Garnet
M-60-80
Garnet
M-36
Description
ES UC
tnm
0.89
1
0
1
1
1
1
1
0
0
0
0
1
0
0
0
0
.38
.98
.08
.16
.30
.46
.92
.54
.60
.72
.75
.20
.60
.25
.17
.47
1.53
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
.12
.43
.53
.49
.47
.38
.72
.41
.28
.42
.43
.08
.07
.10
.71
.43
Initial
depth,
in.
12.75
6.
12.
12.
11.
11.
8.
6.
12.
9.
12.
11.
12.
12.
12.
14.
15.
,38
,75
,25
,12
20
00
63
40
75
38
38
65
35
50
80
45
Downflow Upflow
Rate, Temp, Temp,
gpm/sq ft °C °C
7,8,9 18,22,26,30 18,22,26,30
4,8 18,26 22
4,8 18,26 22
4,8 18 22
4,8 18 22
4,8 18 22
4,8 18 22
4,8 18 22
4,7,8,9 18,22,26,30 18,22,26,30
4,8 18 22
4,8 18 22
4,8 18 22
16,25
16,25
17,25
17,25
17,25
Subscript 2 signifies media remaining after hydraulic grading and skim-
ming  10% of the fines from the raw media of same letter designation.
                                   248

-------
Table 32.  Upflow and downflow experimental runs with dual-media  filters
           in the 6-in. column.
Downflow
Series
B-l
B-2
B-3
B-4
B-5
B-6
B-7
B-8
B-9
B-10
B-ll
B-12
B-13
B-14
Dual
mediaa
AA
AB
AC
AD
A2E
AF
A2A2
AUCA
A2C2
A2F2
A2F2
A2F2
CF
C2F2
Depth, in.
Sand
A =
A =
A «
A =
A2 =
A =
A2 =
A =
A2 =
A2 =
A2 =
A2 =
C =
C2 =
11.25
6.26
10.75
13.25
12.00
6.25
12.00
6.25
9.75
6.63
6.63
6.63
12.25
11.25
Coal
A =
B
C
D
E
F
A2 '
UCA =
C2 -
F2 '
F2 =
F2 -
F
F2 =
12.75
12.63
12.25
12.25
9.00
6.50
11.00
6.25
11.12
7.88
16.00
23.75
12.00
10.87
Rate, Temp,
Total gpm/sq ft °C
23.50
18.50
21.75
24.75
20.50
12.00
22.75
12.00
20.00
12.90
20.00
27.75
22.50
20.38
4,8
4,8
4,8
4,8
4,8
4,8
4,8
4,8
4,8
4,8
4,8
4,8
4,8
4,8
18,26
18,26
18
18
18
18
18
18,26
18
18
18
18
18
18
Upflow
Temp,
°C
22
22
22
22
22
22
22
22
22
22
22
22
22
22
 Described by the letter designation for  the component sand and coal,
 respectively, from Table 31.
                                     249

-------
Table 33.  Upflow experimental  runs with various uniform single media and
           uniform dual media in 2-in.  column, 25 °c.
Series
C-l









C-2



C-3






C-4











D-l












Uniform media description
Sand A (-10+12)
Sand A (-12+14)
Sand A (-14+16)
Sand A (-16+18)
Sand A (-18+20)
Sand A (-20+25)
Sand A (-25+30)
Sand A (-30+35)
Sand A (-35+40)
Sand A (-40+45)
Coal F (-4+7)
Coal F (-7+8)
Coal F (-8+10)
Coal F (-10+12)
Coal A (-10+12)
Coal A (-12+14)
Coal A (-14+16)
Coal A (-16+18)
Coal A (-18+20)
Coal A (-20+25)
Coal A (-25+30)
Coal F (-4+7), 5.9 in. and
Sand A (-40+45), 6.6 in.
Coal F (-4+7), 6.7 in. and
Sand A (-35+40), 5.5 in.
Coal F (-4+7), 5.8 in. and
Sand A (-30+35), 6.0 in.
Coal F (-4+7), 5.7 in. and
Sand A (-25+30), 6.3 in.
Coal F (-4+7), 5.5 in. and
Sand A (-20+25), 5.5 in.
Coal F (-4+7), 5.4 in. and
Sand A (-18+20), 5.8 in.
Garnet (-50+60)
Sand A (-20+25)
Sand A (-30+35)
Sand A (-35+40)
Sand A (-40+45)
Sand A (-20+25), 12.75 and
Garnet (-50+60), 11.40
Sand A (-30+35), a and
Garnet (-50+60), 11.40
Sand A (-35+40), 13.85 and
Garnet (-50+60), 11.40
Sand A (-40+45), 9.80 and
Garnet (-50+60), 11.40
Depth, in.
12.20
12.65
11.90
11.25
11.90
12.00
12.85
11.75
11.65
10.20
12.60
11.95
4.85
13.15
12.05
11.60
11.50
11.20
10.90
10.65
12.50

10.0

10.5

10.0

10.3

9.5

10.3
11.40
12.75

13.85
9.80

23.9

22.25
25 10
fmj • iW
20.80

    aNot recorded.


                                   250

-------
        lOr
     E
     E
      •*.
     UJ
     LIU
    3.0-

    2.0
    1.5-

    1.0
       0.6-
       0.5
       0.4-
       0.3
O.ll	
 0.01  0.1
5
                                               SAND A  V
                                               SAND A2 A
                                               SAND C  O
                                                            I
                    1 2  5^TO  20 30  50     80 90 95

                        PERCENT PASSING BY WEIGHT
                                                   99   99.9  99.99
   5.0
   4.0

   3.0
^ 2.0(-
   1.0
    .9
              Fig. 60.   Sieve analysis of  graded sand media.
     8
8:
0.7
0.4
0.5
0.4
                             COAL A
                             COALB
                             COALC
                             COAL D  o
                             COALE
                             COALF
                                                       A
                                                       O
                                                       COAL C2 D

         '   'I   I	1	I	I    I   I  I   I   I    I     III
     0.1   0.5 1   2   5   10    20  30 40 50 40 70
                                                  90   95  98  99.5
                      PERCENT PASSING §Y WEIGHT

              Fig. 61.   Sieve analysis of graded coal media.
                                  251

-------
          GARNET - 14 4 16 O
          GARNET - 25 + 30 O
          GARNET - 50 4 60 D
          GARNET M-60-80   &
          GARNET M-36      0
    0.
0.01   072
                                                          99.rW.99
                       PERCENT PASSING BY WEIGHT
           Fig. 62.  Sieve analysis of garnet sand media,
 coefficients for the graded sands and coals have been presented pre-
 viously in Table 29.

 In two cases, 10% of a medium was removed by skimming the upper-
 most layer of the filter bed through the use of  a siphoning hose
 after the medium had been stratified by backwashing.   Unskimmed
 graded media were alphabetically designated, for example, Sand A,
 whereas skimmed media were additionally labelled with a subscript,
 for example, Sand A«.  Figures 60 and 61 demonstrate  that a slightly
 larger effective size and a smaller uniformity coefficient resulted
 from skimming as the finest media particles were removed.

 The graded media, as shown in Figs.  60, 61, and  62, followed a near
 log-normal distribution.

 The results of the sieve analyses of the uniform silica sands and
 uniform coals are presented in Appendix A.   The  results indicate
 that at least 70% (generally greater than 80%) of the  total sample
was retained between the adjacent sieves indicated in  the size de-
 signation (e.g., -20+25).
                                 252

-------
Average Particle Size Determinations

The average sizes of the particles for the uniform media were deter-
mined from the sieve analysis presented previously and the equivalent
diameter of a spherical particle.  The data for the equivalent diameter
of a spherical particle is presented in Table 34.  A summary of the
results for the average particle size for the various methods is pre-
sented in Table 35 .  The table includes average diameter by the
arithmetic mean defined by Eq. (39), the inverse diameter defined by
Eq. (38), the diameter of the 60% finer size from the sieve analyses
plotted in Figs. 60, 61, and 62, and the equivalent diameter of a
spherical particle by Eq. (40).

It is apparent from Table 34 that the equivalent spherical diameter
of the various media is larger than the mean of adjacent sieve size
for the uniform media.  This is to be expected because of the non-
sphericity of the media.

Table 35 indicates that the inverse definition for the graded media
gave a somewhat lower average diameter than the arithmetic mean
diameter.  The diameter corresponding to the 60% finer size was
closer to the arithmetic mean diameter.

Densities

The results of the density determinations for the garnet sand, silica
sands, and coal are given in Table 36.  Several replicates are reported
for the garnet sand to illustrate the precision of the analysis.

The garnet  sand used  for density determinations was  from the uniform
-14+16 and  -25+30 sizes  (Table  36).  The density of  the -25+30 garnet
was somewhat  lower than that  of the  -14+16  garnet  sand.  The pycno-
meter bottles were shaken vigorously to remove  air attached  to the
particles  and allowed to cool to  room  temperature.   Therefore, it was
assumed  that  the  inclusion of air or the  temperature effects on  the
density  of water  were not the source of the difference.  The probable
source of difference  was that the -25+30  garnet  contained  a  small
amount of  the less-dense, grayish-colored material that was  assumed
to be removed.  The difference  was  quite  small;  therefore, the average
of all six determinations was used  in  all necessary  calculations in-
volving  garnet sand media.

Porosities

The fixed-bed porosities  for the media used in the experiments were
determined two ways;   (1)  the graduate cylinder technique  and  (2)
 the column technique  (using  both the 2-in.  and 6-in. diameter  fluidi-
 zation columns).   The results for the  porosity determinations  by the
 graduate cylinder technique  and by  the column technique are  given in
Table  37.   In some cases, data were collected by both investigators
 (Woods  [143], Boss [14]).  Both values are  reported  to indicate the
 spread between measurements  made by different investigators.
                                   253

-------
Table 34.  Average diameter of uniform media by  two methods, mean of
           adjacent sieve sizes and mean equivalent spherical diameter
           by the count and weigh method [Eq.  (40)].
                   Particles
   Mean of
adjacent  sieve
Mean equivalent
   spherical
Uniform media
mesh range
Sand A (-10+12)
Sand A (-12+14)
Sand A (-14+16)
Sand A (-16+18)
Sand A (-18+20)
Sand A (-20+25)
Sand A (-25+30)
Sand A (-30+35)
Sand A (-35+40)
Sand A (-40+45)

Coal F (+4)
Coal F (-4+7)
Coal F (-7+8)
Coal F (-8+10)
Coal F (-10+12)
Coal A (-12+14)
Coal A (-14+16)
Coal A (-16+18)
Coal A ( -18+20)

Garnet (-14+16)
Garnet (-14+16)
Garnet (-25+30)
Garnet (-25+30)
Garnet (-50+60)
Garnet (-50+60)

Number
counted
100
100
100
100
100
100
100
100
100
100

100
100
100
100
100
100
100
100
100

110
110
110
110
98
91

Weight,
g
1.0754
0.5752
0.3532
0.2190
0.1514
0.0434
0.0406
0.0231
0.0174
0.0095

12.8546
3.5898
1.8299
1.3413
0.8490
0.3410
0.3040
0.1331
0.0791

0.5798
0.5968
0.0729
0.0685
0.0042
0.0042

sizes, mm
(1)
1.840
1.545
1.300
1.095
0.920
0.775
0.650
0.545
0.460
0.385

—
3.790
2.595
2.180
1.840
1.545
1.300
1.095
0.920

1.300
—
0.650
—
0.273
—

diameter, mm
(2)
1.98
1.61
1.36
1.16
1.02
0.78
0.66
0.55
0.46
0.41
Sand mean
5.22
3.40
2.72
2.46
2.10
1.57
1.51
1.15
0.92
Coal mean
1.347
1.360
0.675
0.661
0.270
0.277
Garnet mean
(2) /(I)
1.08
1.04
1.04
1.06
1.11
1.07
1.02
1.01
1.00
1.06
1.05
—
0.90
1.05
1.13
1.14
1.02
1.16
1.05
1.00
1.08
1.036
1.046
1.039
1.017
0.989
1.015
1.024
Settling Velocities

The experimentally determined settling velocities for  the  three  dif-
ferent uniform garnet  sand media  at  the  two  different  temperatures are
given in Table 38.  The mean time, the standard deviation,  and mean
velocity of the time measurements are included in Table  38.

A comparison of these  experimental values of Vs with the velocity
intercept (Vj_) at a porosity equal to one on the log V vs  log e  plots
                                  254

-------
Table 35.  Summary of the average diameters of the media - d (by
           several methods).
Size
Sand A
Sand C
Coal A
Coal B
Coal C
Coal D
Coal E
Coal F
Garnet (-14+16)
Garnet (-25+30)
Garnet (-50+60)
Garnet (M-60-80)
Garnet (M-36)
Source :
«%•
nun
0.711
1.008
1.285
1.305
1.572
1.988
2.094
3.228
1.275
0.639
0.265
0.272
0.642
Eq. (39)
"inverse'
HKQ
0.685
0.946
1.205
1.246
1.480
1.800
1.912
2.941
1.271
0.637
0.263
0.249
0.615
Eq. (38)
d60% finer,
nun
0.74
1.03
1.35
1.35
1.65
1.90
2.00
3.30
1.27
0.635
0.267
0.29
0.67
Figs. 60, 61, and 62
presented later in Table 41 shows that the experimental Vs was sub-
stantially lower than the Vj^.

These results are consistent with the findings of Carvalho for
crushed coal [28] previously discussed and can be attributed to the
nonspherical shape of the particles.
                                255

-------
 Table 36.  Densities of media — p .
                                  s
Media
Garnet sand (-14+16)
Garnet sand (-25+30)
Silica sand A
Silica sand C
Coal A
Coal B
Coal C
Coal D
Coal E
Coal F
Density of sample, g/ml Densitvf
123 g/ml
4.1337 4.1384 4.1359
4.1228 4.1243 4.1201 Avg 4.13
2.65
2.65
1.70
1.70
1.71
1.72
1.74
1.73
The average settling velocities for the -25+30 and -50+60 garnet sands
were slightly faster at 17 °C than at 25 °C, contrary to expectations.
However, the differences are not statistically significant.
          Fixed Bed Hydraulic  Profiles  in Dual-Media Filters -

                            Coal  and Sand

Downflow Observations  of  Single-Media  Filters

To observe  the  effect  of  various  degrees of intermixing on the per-
meability of dual-media filters,  head  loss vs depth was observed for
the individual  graded  sand and graded  coal media.  These hydraulic
                                  256

-------
Table 37.  Fixed-bed porosities of the three media determined by the
           two techniques (c ) and two investigators.
                            o
                                Porosities
Graduate
cylinder
Media technique
Sand A (-10+12)
Sand A (-12+14)
Sand A (-14+16)
Sand A (-16+18)
Sand A (-18+20)
Sand A (-20+25)
Sand A (-20+25)
Sand A (-25+30)
Sand A (-30+35)
Sand A (-30+35)
Sand A (-35+40)
Sand A (-35+40)
Sand A (-40+45)
Sand A (-40+45)
Garnet sand (-14+16) 0.454
Garnet sand (-25+30) 0.504
Garnet sand (-50+60) 0.550
Garnet sand (M-60-80) 0.557
Garnet sand (M-36) 0.508
Coal F (-4+7)
Coal F (-7+8)
Coal F (-8+10)
Coal F (-10+12)
Coal A (-10+12)
Coal A (-12+14)
Coal A (-14+16)
Coal A (-16+18)
Coal A (-18+20)
Coal A (-20+25)
Coal A (-25+30)
Column Column
technique technique
(2 in.) (6 in.)
0.440
0.442
0.443
0.448
0.440
0.439
0.436
0.444
0.451
0.441
0.456
0.458
0.457
0.452
0.453
0.505
0.554 0.580
0.536
0.495
0.554
0.560
0.573
0.573
0.551
0.555
0.552
0.554
0.554
0.558
0.572
Investigator
*>..}£]
Boss JJ
Boss "
Bo.. "
Bosslri,J01
tt , [143]
Woods
Boss [14]
Boss [14]
Woods']
Boss [14]
Woods [143]
Boss [14]
Boss [14]
Woods [143]
Woods |jJ3j
Woods Jg
Woods J14,J
Woods |"J
Woods [U3J
[14]
Boss|l4j
Boss 4]
Boss £
Boss 14
Boss 141
Boss 14
Boss 14
Boss[l4]
Boss 14]
Boss 4
Boss1 J
profiles for graded media of 6 to 12-in. depth were studied in the
6-in. fluidization column.  The hydraulic profiles of the media
were then to be compared to the hydraulic profiles of the correspond-
ing dual-media filters, comprised of the same components, to observe
the effects of intermixing.
                                 257

-------
 Table 38.  Settling velocities of uniform garnet sand media.
                                   Garnet sand media
                      -14+16               -25+30              -50+60
 Temperature,°C    16.5      25.0      17.0      25.0     17.0     25.0

 Number of
 particles
 observed         100       100       100      100       60       60
          a
 Mean time,
 sec                3.56      3.55      6.92      7.00    17.86     18.53

 Standard
 deviation,
 sec                0.553     0.615     0.993     0.991    2.953    3.151

Mean velocity,
fps                0.703     0.703     0.362     0.357    0.140    0.135

 a
 Distance of timed  free fall = 2.5 ft.
Several variables  affecting head  loss  in a filter are correlated
in Eq. (17) and are evident in Figs. 63  and 64.   The variables have
previously been studied  and correlated in equations  by several in-
vestigators.  However, the figures  are included  here to help the
reader visualize the nature of the  raw downflow  data collected.

A fluid temperature of 18 °C, the lowest consistently attained temper-
ature with the laboratory water system,  and an 8 gpm/sq ft flow rate
were chosen for the downflow measurements.  At high  filtration rates,
the head losses were more substantial  and could  thus be measured with
greater relative precision.  Thus,  the high flow rate of 8 gpm/sq ft
was chosen for the experiments.

Given the same flow rate and temperature, the  difference in cumulative
head losses between two  plotted points in Figs.  63 or 64 is the dif-
ference in pressure between two adjacent piezometers on the large
column.  From this pressure difference,  the experimental values of
head Ipss per 1-1/2-in.  depth were  taken to plot the hydraulic profile.
Bed depth vs the head loss per 1-1/2-in. depth was then plotted in
Figs. 65 and 66 to provide downflow hydraulic  gradient profiles of
all graded media used.   Analysis  of Figs. 65 and 66  demonstrates that
the top sand layer had a head loss  of  0.35 ft, while the bottom coal
layers had a head  loss of 0.01 to 0.04 ft.  Since precision of piezo-
                                  258

-------
vO
               1.00
               o.ao —
              0.40 —
           4:
            *
           r
                          SAND A
                          FLOW RATE 9
                          30.0 °C  O
                          26.5 °C
                          22.0
                          18.0 O
              0.40 —
              0.20 —
                   0.00
0.50       0.75        1.00        1.25
   CUMULATIVE H£Af> LOSS, fc, ft
1.50
1.75
                   Fig. 63.   Fixed-bed  head loss for graded sand A at 9 gpm/sq ft for various
                             temperatures.

-------
              1.00
N>
0>
o
              0.80
   0.60 h~
o
Q
                                                                      SAND A
                                                                      TEMPERATURE

                                                                      9 gpm/*q ft o
                                                                      8 gpnviqft

                                                                                O
                   0.00      0.25        0.50        0.75        1.00        1.25

                                           CUMULATIVE HEAD LOSS, h, ft


                Fig.  64.   Fixed-bed head loss for graded  sand  A at 26.5  C for various  flow rates.

-------
meter readings was limited to 0.01 ft, the small head losses between
adjacent piezometers resulted in larger relative errors in the coal
head loss readings.

Study of Fig. 65, the hydraulic profile of the basic sand, reveals
that a great head loss change occurs in the top half of the filter
bed depth.  Head loss throughout the graded coal media filter, as
shown in the hydraulic profile, Fig. 66, remained fairly constant
with depth.  The surface head loss in the graded coal media filter
was only slightly larger than that of the lower layer.  Therefore,
skimming of the sand media is expected to have more impact on filter
performance than skimming of the coal media.

Downflov Observations of Dual-Media Filters

To determine the effect of intermixing on the fixed-bed hydraulic
profiles of dual-media filters is one of three specific aims of this
research, and Figs. 67 through 72 illustrate typical related experi-
mental findings.  Figure 67 is a typical example of cumulative head
loss vs bed height for dual media M.  The gradual bend of the plotted
lines in Fig. 67 exemplifies the effect of modest intermixing on head
loss.

Since pressure readings could only be made to the nearest 0.01 ft,
some of the data points were relatively inaccurate.  Such points on
figures such as Fig. 67 were adjusted by drawing the best fitting,
curved line for the cumulative head loss.  Values from this curved
line were then chosen at 1-1/2-in. increments of the filter bed.
These values were used to plot the solid-line hydraulic profile of
the dual-media filters, as in Figs. 68 through 72.  The adjusted points
did not vary by more than ± 0.02 ft from the original experimental
points.

Figures 68-72 also provide combined, individual hydraulic profiles
of the dual-media filter components.  These dashed-line hydraulic
profiles show the head loss which would result at the interface if
intermixing did not occur.  To obtain these theoretical profiles, the
hydraulic profiles for the appropriate sand and coal from Figs. 65
and 66 were combined.  Coal was plotted above sand, and the two
hydraulic profiles were connected with a horizontal line.  The hori-
zontal connecting line was plotted at the bed height corresponding
to the sand depth indicated in Table 32.  Thus, the hydraulic profiles
of a hypothetical nonintermixed filter and actual intermixed filter
are shown in Figs. 68 through 72.  In Figs. 69 and 72, where the
sand or coal depth was not the same as in Figs. 65 and 66, the gradient
curve for the sand and coal was compressed vertically to fit the ex-
perimental sand depth.  The actual observed location of the inter-
mixing zone is shown in Figs. 68 through 72.  The upper limit of
the intermixing zone corresponds to the lowest filter level where only
coal particles were evident, and the lower limit corresponds to the
                                  261

-------
              1.00
NJ
0*
NJ
          £
          LU
          Q
              0.80
              0.60
0.40
              0.20
              0.00
GRADED SAND
FLOW RATE 8 gpm/«q ft
TEMPERATURE 18 °C
SAND A  O
SAND A2 A
SAND C  D
SAND C2 •
                  0.00       0.05        0.10       0.15       0.20        0.25        0.30
                                 HEAD LOSS PER 1-1/2-m. DEPTH, PLOTTED AT TOP OF THE
                                                DEPTH INCREMENTS, ft
           Fig.  65.  Head loss for individual graded media in 1-1/2-in. unit filter sections.
                                                                              0.35

-------
CO
                                                       COM.
                                                       COAL
                                                       COAL
                                                       COAL
                                                       COAL
                                                       COAL
                                                       COAL
                                                       COAL
                                                       COAL
                                                       COAL
       0.00     0.01        0.02       0.03       0.04       0.05        0.06
                   HEAD LOSS PER 1-1/2-in.  DEPTH, PLOTTED AT TOP OF THE
                                DEPTH INCREMENTS, ft
Fig.  66.  Head loss for individual graded media in 1-1/2-in. unit filter sections.
                                                                                           0.07

-------
              2.00
          *

            %


          JE

          LU

          Q

          Q
to
^
4s
                                                                            DUAL MEDIA AA

                                                                               o
                                             18"C4gpin/sqft  O
              0.50 -
              0.00
                   0.00
0.25
0.50
0.75
1.25
                                             CUMULATIVE HEAD LOSS, ft
                 Pig. 67.  Fixed-bed head loss of dual Media AA at various temperatures and flow rates.

-------
10
£
X
LU
Q

2
CD

85
                    2.00
                    1.60
                    1.20
                    0.80
                    0.40 —
                    0.00
                        0.00
                                        TEMPERATURE 18 °C
                                        FLOW RATE 8 gpm/iq ft

                                         INTERFACE RANGE
                                         UPPER LIMIT     T~
                                         AVERAGE LIMIT	
                                         LOWER LIMIT
0.05
                             0.10
0.15
0.20
0.25
0.30
0.35
                                        HEAD LOSS PER 1-1/2-in. DEPTH, PLOTTED AT TOP OF THE
                                                     DEPTH  INCREMENTS, ft
                Fig. 68.  Head  loss  per 1-1/2-in.  unit depth in dual media AA and head loss for the
                          two-component media if unmixed.

-------
NJ
                   2.00
                   1.60
                   1.20
                04
                Q
                   0.80
                   0.40
                   0.00
                       0.00
                                   COAL C
                        DUAL MEDIA AC
                                        INTERFACE RANGE
                                        UPPER LIMIT    ~T
                                        AVERAGE LIMIT —
                                        LOWER LIMIT   _L
                                                                         TEMPERATURE 18 °C
                                                                         FLOW RATE 8 gpm/sq ft
                                                                  I
                                             I
0.05
0.10
0.15
0.20
0.25
0.30
0.35
                                    HEAD LOSS PER 1-1/2-in. DEPTH, PLOTTED AT TOP OF THE
                                                 DEPTH  INCREMENTS, ft
               Fig. 69.   Head loss per 1-1/2-in. unit depth in dual media AC and head loss for the
                         two-component media if unmixed.

-------
     2.00
     1.60
 
-------
N>
ON
00
                   2.00
                   1.60
                   1.20
x
£  o.so
Q
9
oo
55
5  0.40
                   0.00
                                                           INTERFACE RANGE
                                                           UPPER LIMIT
                                                           AVERAGE LIMIT
                                                           LOWER LIMIT
                                                                         TEMPERATURE 18°C
                                                                         FLOW RATE 8 gpm/sq ft
                       0.00
                 0.05
0.10
0.15
0.20
0.25
0.30
0.35
                                        HEAD LOSS PER 1-1/2-in. DEPTH, PLOTTED AT TOP OF THE
                                                     DEPTH INCREMENTS, ft
              Fig. 71.  Head loss per 1-1/2-in. unit  depth in dual media A~E and head loss for the
                        two-component media if unmixed.

-------
a*
VO
1.60

^
£ 1.20
t
LU
0
a
CD
2j 0.80
^

0.40



0.00
0.





/ M T
1 V
-1- COAL F ^s.
L 	 	




1 1
00 0.05 0.10

INTERFACE RANGE
UPPER LIMIT T~
AVERAGE LIMIT —
LOWER LIMIT J_


DUAL MEDIA AF
^L_ -----jr^-^^-^ 	
^^-a^- — . — — — — ^^
TEMPERATURE 18°C
FLOW RATE 8 gpm/sq ft
SAND A
1 1 1 1
0.15 0.20 0.25 0.30 0.












35
                                  HEAD LOSS PER 1-1/2-in. DEPTH,  PLOTTED AT TOP OF THE
                                               DEPTH INCREMENTS, ft
                 Fig.  72.   Head loss per 1-1/2-in.  unit depth in dual media AF and  head  loss for the
                           two-component media if unmixed.

-------
 highest filter level where only sand particles were evident.   The
 average interface was judged as the filter level at which the coal
 and sand particles are approximately equal in number.   Of course,
 these intermixing zone distinctions are based on visual observations
 and are subject to errors in judgment.

 Figure 68 shows a typical example of head loss reduction at the inter-
 face, which results when intermixing occurs.   Without  intermixing, the
 head loss gradient in the sand layer at the interface, as shown in
 Figure 68, would be expected to be 0.35 ft.  Intermixing of the coal
 and sand reduced this head loss gradient to 0.17 ft, a 50% reduction
 in the head loss.   This reduction was attributed to intermixing,  be-
 cause the zone of intermixing was visually observed to occur  when-
 ever the experimental hydraulic profile differed from  the individual
 component hydraulic profiles.  As the very fine sand mixed up into
 the coarser coal layer, a redistribution of the hydraulic profile
 was observed,  causing a greater head loss in  the coarser coal layer,
 with a correspondingly decreased head loss in the sand layer.  The
 enclosed areas of Figs. 68 through 72,  which  illustrate this  decrease
 in head loss,  labelled I,  in the sand layer and increase in head  loss,
 labelled II, in the coal layer,  are approximately equal in area,  ex-
 cept in cases  where excessive intermixing occurred.

 The installation of various  graded coal media with the same graded
 sand medium, as illustrated  in Figs.  68 through 70,  did not signi-
 ficantly change the 0.17>~ft  maximum head loss at the interface.
 Therefore,  the graded sand medium was deemed  as controlling the
 interface head loss,  regardless  of the  coal media involved in the
 intermixing process.   The  effect of using increasingly coarser coals
 with the same  sand is also evident in Figs. 70 through 72.  The
 portion of  bed showing intermixing increases  with the  coarser coals.
 The general effect of intermixing in  Figs.  68  through  72 is to cause
 a  gradual decrease in permeability with depth.   This decrease would be
 due to  decreased  average pore dimension and decreased  porosity.  The
 influences  of  these variables on head loss  (permeability)  are evident
 in Eq.  (17).

 The general effect of intermixing is  to provide  gradual  coarse to
 fine  filtration in the  direction of flow.  The  desirability of this
 coarse  to fine  filtration  is  accepted by all  researchers  studying
 filter performance.  The fact that the  peak hydraulic  gradient of
 the dual-media  filter  is less  than the  peak gradient of  the sand
 filter alone would indicate  that  the  effectiveness of  the sand layer
 in  filtration would be  somewhat diminished.  However,  the diminishment
would be offset by the  improved  effectiveness of  the lower coal layers
where intermixing  is present.  The  prediction of  relative filter per-
 formance, in light of  these observations,  cannot be made  at this  time.

Figures  70  through 72 demonstrate  excessive intermixing.  Sand
particles actually reached the surface  of  the coal layer  and  caused
                                 270

-------
an increase in the head loss throughout the coal layer as the coal
bed permeability decreased due to the presence of the sand.  The head
loss increase in the coal layer was especially significant, as shown
in Fig. 72.  The capacity of the dual-media filter could be expected
to be decreased due to the finer pore size in the surface layers of
the filter.  Thus, Fig. 72 gives an example of too much intermixing.

The water shutdown procedure following backwashing was also considered
important to downflow hydraulic profiles and thus to filter performance,
Preliminary work was done with dual media A2C2 to determine the effect
of shutdown on intermixing.  Two shutdown procedures, a slow 1-min
valve closure and a fast 5-sec valve closure, were tested.  The bed
height was 12.9 in. for all test measurements.  Given high fluidizing
flow rates, a large amount of intermixing was achieved before shut-
down.  A fast shutdown procedure allowed for a greater amount of the
intermixing to be retained as the coal and sand particles settled
than did a slower shutdown procedure.  The slower, 1-min procedure
allowed the particles of the two media to separate and stratify.

                  Expansion — Flow Rate Observations

Garnet Sand

The raw data collected during expansion of garnet sands in the 6-in.
column were;  (1) flow rate, (2) bed height,  (3) manometer readings
at 3-in. increments of the bed, and (4) temperature.

The following analyses of these data were made.

1.  Flow meter reading was corrected by the  appropriate calibration
    equation and  changed to velocity (in gpm/sq  ft and fps) based
    on the open cross-sectional area of the  column.

2.  Bed height reading was corrected by adding 1 in.  The  average
    porosity ratio  of the expanded bed was calculated from the
    expanded height of the bed by Eq.  (2) using  the previously
    determined e   and the observed 1 .
                o                  o

3.  Pressure loss through the bed was  calculated by  subtracting
     a  piezometer  reading above the expanded  bed  from the piezometer
     reading  at the  bottom of  the bed column.

A typical  example of  the data  (run 1,  Series A-13) is given  in
Table  39.

The  computed values of  the  data  for  each  expansion run were  plotted
as pressure  vs  flow rate,  e.g., Figs.  73  and 74, and expanded bed
height vs  flow  rate,  e.g.,  Figures  75  through 84.

An analysis  of  the pressure loss vs  flow  rate figures leads  to the
following  observations:
                                   271

-------
Table 39.  Expansion — flow rate data of run 1, Series A-13  (-14+16
           garnet sand media).

Flow rate
gpm/sq ft
110.0
99.8
93.0
82.3
73.7
68.2
61.3
56.4
51.8
48.8
45.4
42.6
38.6
35.4
32.5
27.5
23.6
18.6
12.2
Corrected
(V)
fps
0.246
0.222
0.207
0.184
0.164
0.152
0.137
0.127
0.116
0.109
0.101
0.0950
0.0859
0.0788
0.0725
0.0613
0.0525
0.0415
0.0273

Bed height
U),
in.
20.50
19.50
18.75
18.00
17.75
17.25
16.63
16.25
15.75
15.63
15.50
15.25
14.88
14.50
14.25
13.80
13.70
13.65
13.65
Average
porosity
ratio
(e)
0.636
0.618
0.603
0.586
0.580
0.568
0.552
0.541
0.527
0.523
0.519
0.511
0.499
0.486
0.477
0.460
0.456
0.454
0.454
Pressure loss
through bed,
ft
1.83
1.87
1.87
1.90
1.90
1.90
1.88
1.88
1.87
1.86
1.86
1.85
1.85
1.86
1.86
1.73
1.41
1.04
0.64
      for  this run collected during contracting  of  fluidized bed  only,
 Water  temperature =  16.5°C.
                                  272

-------
  2.0
               RUN 1

               TEMPERATURE 16.5 °C



000 0000 OOOOO    00   O
   1.0
c  0.0.
8
UJ
at

to
e
   2.0
               0.05
      0.10
0.15
0.20
                                     RUN 2
                                     TEMPERATURE 25.0 °C
                         OOOQO oo o  o
                        O   O   O
0.25
   1.0
   0.0
     0.0
      OJO       0.15

    FLOW RATE, V,  fps
            0.20
            0.25
    Fig.  73.  Pressure loss - flow rate diagram for garnet sand

             media (-14+16).
                               273

-------
   2.0
   1.0
c o.o ^
 „   0.0
8
LU
  2.0|
  1.1
                                       RUN 7
                                       TEMPERATURE 17.0 °C
                «••• •
OT02
0.04
               0.06
              0.08
RUN 8
TEMPERATURE 25.0 °C
   OA4        0.06
FLOW RATE, V, fps
             0.08
                                                               0.10
                                                               0:10
  Fig.  74.   Pressure loss - flow rate diagram  for garnet sand
            media  (M-60-80).
                                274

-------
                                 WEN AND YITS  MOOR
                                 AUTHOrS MCCB. fe
to
•vl
Ui
               If
              17 —
              13
                                O O
                                                                         MEDIA - 14 + 16
                                                                         TEMPEIATUKE 16 °C
                          0.04
0.08
0.20
                             0.12        0.16

                            FLOW RATE, V, fps

Fig.  75.   Expansion - flow rate characteristics  (garnet sand, run 1).
0.24
0.28

-------
K>
••4
              21
          —  17
          m  •'
          x
              15
              13
                                 WEN AND YU'S MODa
                                 AUTHOR'S MODEL 60
                       O   O  O
                          0.04
                                    MEDIA -14 + 16
                                    TEMPERATURE 25.0 °C
0.08
0.20
                             0.12        0.16

                           FLOW RATE, V, fps

Fig.  76.  Expansion - flow rate characteristics (garnet sand, run 2).
0.24
0.28

-------
ro
                                      WEN AND YU'S MODEL
                                      AUTHOR'S MODEL 6a
                                                                          MEDIA -25+30
                                                                          TEMPERATURE 16.0 °C
                                       0.04       0.06
                                               FLOW RATE, V, fps
0.10
0.12
0.14
                    Fig. 77.  Expansion - flow rate characteristics (garnet sand, run 3).

-------
NJ
~J
oo
               23
             ,  19
O
UJ


a
               15
               13
                                  WEN AND YU'S MODEL

                                  AUTHOR'S MODEL 60
                                                                                           0
                                                                                       tf
                        00,0 00)
                                                                           X
                                                                  X
                                                               MEDIA -25+30

                                                   X         TEMPERATURE 25.0 °C
                                                          X
                           0.02
                           0.04
0.10
                              0.06        0.08

                           FLOW RATE, V, fps

Fig.  78.  Expansion - flow rate characteristics (garnet sand, run 4).
0.12
O.U

-------
   40
   38 —


   36


   34


   32


   30
!E
o
26


24

22


20


18

16

14
             WEN AND YU'S MODEL   /
             AUTHOR'S MODEL 60      /
             AUTHOR'S MODEL 6b     /
                                      MEDIA -50 + 60
                                      TEMPERATURE 17.0 °C
               0.02        0.04       0.06
                        FLOW RATE, V, fps
                                              0.08
0.10
  Fig. 79.   Expansion - flow rate characteristics (garnet sand,
            run 5).
                              279

-------
                  WEN AND  YU'S MODEL
                  AUTHOR'S MODEL 6a
        	AUTHOR'S MODEL 6b
                                      MEDIA -50 + 60
                                      TEMPERATURE 25.0 °C
             0.02       0.04        0.06
                      FLOW RATE, V, fps
Fig.  80.  Expansion - flow rate characteristics  (garnet sand,
         run 6).
                            280

-------
            WEN AND YU'S MODEL  /
            AUTHOR'S MODEL 60
          - AUTHOR'S MODEL
                                      MEDIA M-60-80
                                      TEMPERATURE 17.0 °C
                        FLOW RATE,  V, fps
Fig.  81.  Expansion - flow rate characteristics (garnet sand,
         run 7).
                             281

-------
            	WEN AND YU'S MODEL   /
            	AUTHOR'S MODEL 6a    /
           	AUTHOR'S MODEL 6b    6
                                     MEDIA M-60-80
                                     TEMPERATURE 25.0 °C
              0.02
 0.04        0.06
FLOW RATE,  V, fps
0.08
0.10
Fig.  82.  Expansion - flow rate characteristics (garnet sand,
         run 8).
                             282

-------
to
00
   32i



   30



   28
            , 26
g  24
           S
           CO
              16
   22



   20 —



   18
                               WEN AND YU'S MODEL

                               AUTHOR'S MODEL 60
                                                                                             O
                                                                         MEDIA M-36

                                                                         TEMPERATURE 17.0 °C
                          0.02
                          0.04
0.06        0.08

FLOW RATE, V, fps
0.10
0.12
0.14
                    Fig. 83.  Expansion - flow rate characteristics (garnet  sand, run 9).

-------
   32

   30


.E28

^26
 *

§24

   22
LU
r
   18

   16
               	WEN AND YU'S MODEL
                        AUTHOR'S MODEL 60
   20 I                         rv
                               f^             '" MEDIA M-36
                                               TEMPERATURE 25.0 °C
              doo  oPi        r^     I         I
      0      0.02     0.04     0.06     0.08     0.10     0.12   0.14

                            FLOW RATE,  V, fps
   Fig. 84.  Expansion - flow rate characteristics  (garnet sand,
             run 10).
1.  The apparent minimum fluidization can readily be determined
    as the intersection of the two linear portions of the curve.
    These values as well as the Vmf values determined from Wen
    and Yu's equation [Eq. (53)], Leva's equations [Eqs.  (51)  and
    (52)], and Frantz's equation [Eq. (54)]  are listed in Table 40.

2.  The effect of channeling is greater at about Vp and  decreases
    as expansion increases.  The evidence of channeling is the
    lower pressure drop observed for flow values just above minimum
    fluidization.  Channeling is especially noticeable for the finer
    media (Figs. 77 through 82) and was visually observed  and noted
    during the actual fluidization.  The channeling is attributed
    to poor flow distribution into the expansion column.

The bed height vs flow rate figures (Figs.  75 through 84)  also
present the results of calculated bed height for garnet sand media
by various expansion models.   The models and discussion will be
presented later.
                                 284

-------
Table 40.  Summary of minimum fluidization velocities of garnet  sand
           media - V
                    mf
                        Minimum fluidization velocities, fps
                From head loss     Wen and
     Garnet   vs flow rate,  e.g.,    Yu's         Leva's        Frantz's
Run  media     Figs.  73 and 74    Eq. (53)   Eqs.  (51) and (52)   Eq.  (54)
1
2
3
4
5
6
7
8
9
10
-14+16
-14+16
-25+30
-25+30
-50+60
-50+60
M-60-80
M-60-80
M-36
M-36
0.
0.
0.
0.
0.
0.
0.
0.
0.
067
074
027
033
0071
0078
0067
0076
031
0.036
0
0
0
0
0
0
0
0
0
0
.067
.075
.021
.026
.0040
.0048
.0042
.0051
.022
.026
0
0
0
0
0
0
0
0
0
0
.0059
.064
.024
.029
.0049
.0058
.0051
.0061
.025
.029
0
0
0
0
0
0
0
0
0
.158
.196
.040
.049
.0070
.0085
.0074
.0089
.041
0.049
The  analyses  of the  log V vs log e plots for the various media are
the  most important analyses of the expansion-flow rate experiments.

One  typical plot,  log V vs log e (run 1, Series A-13), is shown on
Fig. 85.  The important characteristics of this figure are the
slope of the  line  or n slope and the velocity intercept at porosity
ratio equal to one.   To remove the bias of fitting a straight line
to this plot, a linear regression analysis was performed.  The V
and  e data that were used for the linear regression analysis were
all  of the data points above the intersection of the two straight line
portions of the pressure loss vs flow rate plot, e.g., Figs. 73 and
74.   The results of this analysis for all the expansion flow rate runs
 on garnet sand are given in Table 41.  Included in this table are V±
 in  gpm/sq ft  and  fps, n  slope, number of points used in the regression
 analysis, the correlation coefficient of the log-log line, the
 Reynold's number based on the arithmetic mean diameter and Vt, and
 Richardson and Zaki's n  slope calculated from Eq. (30), neglecting
 d/D.
                                   285

-------
       -0.4
   Fig. 85.
       -0.3        -0.2        -0.1
          LOG POROSITY RATIO, c
Log plot of V vs e for garnet sand media (-14+16)
(run 1, Series A-13.
The experimentally determined n slope values are higher than the n
slope values from Richardson and Zaki's Eq. (30).  This was expected
from the literature and is attributed to the particle shape.  Because
the garnet sand media has a very irregular shape, no attempt will be
made to correlate the n slopes for the data with the calculated n
slopes by Richardson and Zaki's equations.  Rather, a unique equation
for the solution of n slope for garnet sand will be proposed.
                                 286

-------
Following the same approach as Richardson and Zaki, log n vs log
Re^ was plotted (Fig. 86).  The linear regression correlation
equation was determined using all the 10 sets of points from column
1 and column 7 of Table 41.  The resulting equation is,
     n = 5.758 Re
                 -0.0541
                (69)
with a coefficient of determination for the log-log line, r  = 73.32%.

For the evaluation of n slope, the Re. must be defined.  The prop-
erties of the water and the linear dimensions of particle size are
readily available or can be determined.  The available methods for
evaluating the velocity intercept, which equals the settling velocity
of a discrete particle for spherical particles, are discussed in
the section on prediction of settling velocities, beginning on page 229,
    1.00
    0.75
 UJ
 Q.
 2
 to
 o
 Q
    0.50
    0.25
o  o
                            I
  I
        0        0.5       1.0       1.5       2.0       2.5      3.0
      LOG  REYNOLD'S NUMBER, Re. (BASED ON VELOCITY INTERCEPT, V.)
   Fig. 86.  Log plot of n slope vs Reynold's number  -  Re^  (for
             garnet sand media, runs 1 through  10,  Series A-13
             through A-17).
                                   287

-------
 The weaknesses in three methods for determining settling velocity are
 discussed as follows:

 A comparison of the experimental settling velocities of discrete
 particles (Table 38) with V^ (Table 41) does not show any reasonable
 agreement.  This observation is consistent with expectations from
 the literature as previously described (pages 216 through 228) .

 Graphical correlations of the ratio of Re^Re^ vs Ga have the in-
 herent weakness of requiring a family of curves for varying e^ ,
 and the graphs presented did not fit the data points presented in
 the literature as well as would have been desired.

 The graphical solution from a log- log plot of ReQ vs Ga would be
 a fast and simple method of determining settling velocities (pages 229-
 234   ) .   An attempt was made to use the correlation of log Re0 vs
 2/3 Ga of Coulson and Richardson [33, p. 147].  However, values
 of Vg thus determined did not compare well with the experimental
 V^ values in Table 41.  But following the same method of plotting
 the log Re^ vs log Ga (dropping the constant 2/3), the experimental
 values plotted closely as a straight: line (Fig. 87).  The linear
 regression equation of this line is,

      R6i  = 0.0702 Ga0'823                                     (70)
                                                            2
 with a coefficient of determination for the log- log line, r  = 99.38%.
 The values used for this determination and the resulting values  from
 use of Eq. (70) are listed in Table 42.  The agreement of the Re.
 values from the above equation  when compared with the experimental
 Re.  result in an  error of - 14.7 to + 16.7%.


Presentation of Garnet Sand Expansion Model

The modified model for the  expansion of garnet  sand  is proposed based
on  the empirical -correlations presented previously [Eqs.  (69)  and  (70)]
It is outlined  in the following  steps:
1.  Experimentally determine arithmetic mean diameter  (djj), the
    density of the particles (ps), and the fixed-bed porosity  (e f)

2.  Calculate Ga from the fluid and particle properties.

3.  With Ga, calculate Re. from Eq. (70).

4.  Calculate V. from Re  and n slope from Eq. (69).

5.  The porosity at any desired flow rate can then be determined
    from Vi, n slope, and Eq. (28),
                                  288

-------
       Table 41.  Results  of  the log V vs log e relationship for garnet sand media.
ts>
00
Run
1
2
3
4
5
6
7
8
9
10
n
slope
4.326
4.089
4.437
4.420
5.321
5.197
4.932
4.926
4.139
4.173
Velocity
log VL
2.90380
2.87682
2.42580
2.47664
1.8961
1.94649
1.81071
1.87366
2.33674
2.39184
intercept
gpm/sq ft
801.3
753.0
266.6
299.7
78.73
88.41
64.67
74.76
217.1
246.5

-------
  o
  o
     4.0
      3.0
   0>
  a:
  eo


  Z
  Q 2.0
  O
  Z
  fc
1.0
                 1.0       2.0       3.0       4.0
                          LOG GALILEO NUMBER, Go
                                                    5.0
6.0
   Fig. 87.  Log plot of Reynold's number, Re^ vs Galileo number,
             Ga (for garnet sand media, runs 1 through 10, Series
             A-13 through A-17).
      e =
               1/n
6a.  Using the experimentally determined value of e ~ and selecting
     any desired initial bed height (£Q), the expanded height of
     the bed can be determined from Eq. (2),
      I =
           (1 - e)
6b.  As an alternative approach, which eliminates the need to mea-
     sure e^f in step 1, the Vmf can be calculated from a minimum
     fluidization velocity equation.  For this study, Leva's
     equation [Eq. (51)] with Zabrodsky's correction factor [Eq. (52)]
     for Re   > 10 gave the most consistent results (Table 40) with the
                                  290

-------
Table 42.  Values of Reynold's numbers and Galileo's number for the
           garnet sand media.
Run
and
series
1,
2,
3,
4,
5,
6,
7,
8,
9,
10,
A-13
A-13
A-14
A-14
A-15
A-15
A-16
A-16
A-17
A-17
Garnet
media
-14+16
-14+16
-25+30
-25+30
-50+60
-50+60
M-60-80
M-60-80
M-36
M-36
Ga
51,876
80,370
6,530
10,117
490
722
530
780
6,969
10,261
Re^ based
on V^ and
d_
m
625
730
104
146
13.1
17.8
11.0
15.5
87.5
120
Re^ from
Eq. (70)
533
764
96.8
139
11.5
15.8
12.3
16.9
102
140
7, error
- 14.7
4.6
- 6.9
- 4.8
- 12.2
- 11.2
11.8
9.0
16.6
16.7
    V £ values determined from the intersection of  the  two  straight
    l?ne portions of the head loss vs  flow  rate plots.  The emf is
    then calculated from Eq.  (28),

             /.v  \ 1/n
       €mf = (~vf )


    using the V^ and n  slope  from step 4 above.   The  expanded height
    is then calculated  in the same way with Eq.  (2).

The values of Re.^, n, and V^, calculated from Eqs.  (70) and (69)
and the definition of Reynold's  number, are listed  in Table 43.
The emf values calculated from Leva's  Vmf values  of step 6b are  also
listed in Table 43.

The results of expanded bed height calculated using the above model,
including step 6a (designated authors' model  6a), are presented  in
Figs. 75 through 84 for runs  1 through 10.   Similar results for  the
                                  291

-------
 Table 43.   V,,  n. «  . to be used in author's expansion models.
             l       tnr
Run
1
2
3
4
5
6
7
8
9
10
Garnet
media
-14+16
-14+16
-25+30
-25+30
-50+60
-50+60
M-60-80
M-60-80
M-36
M-36
Ga
51,876
80,371
6,530
10,117
490
722
530
780
6,969
10,260
Re^ from
Eq. (70)
533
764
96.8
139
11.5
15.8
12.3
16.9
102
140
n from
Eq. (69)
4.10
4.02
4.50
4.41
5.05
4.96
5.03
4.95
4.49
4.41
Vi (fps)
from Rei
definition
1.52
1.77
0.552
0.637
0.154
0.175
0.160
0.182
0.565
0.641
'./
0.453
0.443
0.497
0.496
0.505
0.504
0.505
0.503
0.497
0.496
 Calculated from emf = (Vmf/Vi)    where V^ determined by Leva's
  equations  [Eqs. (51) and (52)] and V^ and n from  this table.

model using step 6b  (designated  authors'  model 6b)  are presented for
runs 5, 6,  7,  and  8.  The results  of using step  6b  for runs 1 through
4,  9, and  10 do not  differ more  than 1/4  in.  from the  results using
step 6a and, therefore,  are not  shown in  the  figures.

The garnet  expansion model using step 6b  requires a minimum of experi-
mental work.   The  average diameter from the sieve analysis and a
density determination for the particles are all  that is required ex-
perimentally .

For comparison purposes,  the expansion equation  of  Wen and Yu
[Eq. (36)] was  also used  to calculate expanded bed  height, and the
results are presented in  Figs. 75  through 84.  It is apparent that
the Wen and Yu  equation does not provide  good  predicted expansions.

Silica Sand and Coal

The expansion vs flow rate data  for  the uniform  silica sands  and
coals were  analyzed in the same manner  as  the  garnet sands.   This
                                  292

-------
was done  to develop a model which could  be  used to predict expansion
of silica sand or coal using the same  general approach previously
described for garnet.

The expansion vs flow rate data for the  uniform media were first
analyzed  to determine the porosity at  each  flow rate.  Log porosity
was then  plotted against log superficial velocity such as Fig. 85
presented previously.  The slope of these curves n and the inter-
cept velocity (Vi) at e = 1.0 were determined by regression analysis,
The results are presented in Table 44.

The relationship between the n slope and Reynold's number based on
Vi was then determined for all the sand  data  and separately for all


Table 44.   Results of log V vs log e regression  analyses  for uniform
            sized silica sands and coals.
   Media
designation
  n
slope
Velocity
intercept
 Vi,(fps)
Number of
 values
  used
Coefficient
     of
determination
    r2, 7.
based on
V and da
Galileo
 no.
Silica Sand

A-10+12
A-12+14
A-14+16
A-16+18
A-18+20
A-20+25
A-25+30
A-30+35
A-35+40
A-40+45

Coal
2.573
2.665
2.900
2.909
3.153
3.296
3.504
3.704
4.126
4.206
0.557
0.500
0.492
0.420
0.383
0.320
0.281
0.258
0.236
0.187
                   11
                   13
                   13
                   14
                   15
                   16
                   14
                   13
                   14
                   11
                        99.76
                        99.95
                        99.75
                        99.82
                        99.88
                        99.37
                        99.47
                        99.54
                        99.47
                        99.55
347
263
217
157
119
84.7
61.9
47.9
36.9
26.1
125,296
74,939
44,054
26,803
15,662
9,553
5,539
3,327
1,958
1,404
F-4+7
F-7+8
F-8+10
F-10+12
A-10+12
A-12+14
A-14+16
A-16+18
A-18+20
A-20+25
A-25+30
3.016
2.832
2.920
2.939
3.254
3.448
3.667
3.479
3.523
3.988
3.739
ad * arithmetic mean
0.565
0.506
0.465
0.388
0.321
0.291
0.271
0.220
0.186
0.171
0.139
of adjacent
7
7
9
10
11
12
9
10
10
8
7
sieve
99.49
99.68
99.53
99.27
99.64
99.79
99.85
99.86
99.68
99.67
99.43
sizes.
726
447
344
242
201
153
119
82.0
58.1
45.1
30.6

484,270
157,652
92,145
55,558
53,284
31,869
18,734
11,398
6,660
4,062
2,356

                                   293

-------
 the coal data in Table 44.  The results of the regression analysis
 yielded the following relationships:

 For silica sand A, all sizes in Table 44,


      n = 7.973 Re^0'1947                                (71)

                                     2
 with coefficient of determination, r  = 98.16%.

 For coal F and coal A, all sizes in Table 44,

      n = 5.517 Re.'0'1015                                (72)

                                     2
 with coefficient of determination, r  = 78.36%.

 Following the approach of the garnet sand expansion model,  the rela-
 tionships between Re.  and the corresponding Galileo number  (Ga) were
 determined for the uniform silica sands and uniform coals.   The values
 of Ga  are presented in Table 44.   The regression analysis of log Re.
 vs log Ga gave the following relationships:

 For silica sand  A,  all sizes in Table 44,

     Re± = 0.5321 Ga°'5554                               (73)

                                     2
 with coefficient of determination,  r  = 99.74%.

 For coals  A and  F,  all sizes in Table 44,


     Re± =  0.2723 Ga°*6133                               (74)

                                     2
 with coefficient  of determination,  r  = 99.39%.

 The above relationships were then  tested for  validity by  calculating
 the expansion vs  flow  rate data for all  uniform sands and coals.
 The procedure was exactly  as described  previously for garnet sand
 (authors' model 6a) using  the appropriate n vs  Re.  and Re^ vs Ga
 relationship for  the media under analysis.

The results for the prediction  of uniform sand  expansions were  very
 good as expected  from the  high  coefficients of determination for
Eqs. (71) and (73).  The predicted  values of  n  and V± for the various
 sands are presented in Table 45, along with the maximum percent  error
 in the predicted expanded  bed depth up  to a total expansion  of  70%.
 It is apparent from Table  45 that the uniform sand expansions can be
predicted acceptably, with errors not exceeding about 7%  of  actual
 observed values.
                                 294

-------
Table 45.   Predicted values  of n,  V^ with  errors of prediction and
           maximum error pf  prediction  of  expanded bed depth  for
           uniform sands.
Sand
designation
A(-10+12)
AC-12+14)
A (-14+16)
A(-16+18)
A(-18+20)
A(-20+25)
A(-25+30)
A(-30+35)
A(-35+40)
A( -40+45)
Vi
from
Eq. (73),
fps
0.559
0.499
0.443
0.397
0.352
0.316
0.280
0.250
0.222
0.206
%
error
in
V
- 3.64
0.25
'9.51
5.38
7.17
8.42
1.94
0.94
3.60
- 0.45
n
from
Eq. (71)
2.533
2.678
2.836
2.993
3.172
3.346
3.549
3.751
3.972
4.117
%
error
in
na
1.52
- 0.51
2.19
- 2.90
- 0.60
- 1.53
- 1.29
- 1.26
3.73
2.11
Maximum
7, error in
expansion'3
+ 2.2
- 1.1
- 7.5
- 7.1
- 6.5
- 2.3
- 2.1
- 3.3
- 3.6
+ 5.9
 aCompared  to values  in Table 44.

  [(Observed depth -  predicted  depth)  100/observed  depth].  Maximum
  observed  error  in predicted expansion up  to  a  total  bed expansion
  of  70% above  fixed  bed  depth.
On the  other hand,  the  prediction of  coal  expansions were not  as  good
as shown in Table 46.   It  appears that the model  provides an acceptable
prediction for the  uniform sizes  from coal A,  but is unable to do as
well for the uniform sizes from coal  F.  This  may be due, in part,  to
different sphericity for the two  coals.  It may also be  due to less
well-defined uniform sizes due to the angularity  of the  crushed coals
and resulting  sieving difficulties.   More  work should  be done  on  the
crushed coals  to attempt to improve the  prediction accuracy.

The difficulties encountered with the coal expansion prediction empha-
size one important  weakness in the expansion models:   they do  not
incorporate any direct  measure of sphericity.   They are  empirical
                                   295

-------
 Table 46.   Predicted values  of  n,  V^ with errors of prediction  and
            maximum error of  prediction of expanded bed height for
            uniform coals.
Coal
designation
F-4+7
F-7+8
F-8+10
F-10+12
A-10+12
A- 12+14
A-14+16
A-16+18
A-18+20
A- 20+25
A-25+30
Vi
from
Eq. (74)
fps
0.629
0.459
0.395
0.343
0.334
0.289
0.249
0.217
0.187
0.163
0.140
%
error
in Via
- 11.10
9.20
15.13
11.69
- 3.96
0.39
7.85
1.28
- 0.19
4.73
- 0.62
n
from
Eq. (72)
2.787
2.989
3.090
3.189
3.197
3.301
3.413
3.520
3.639
3.753
3.883
7.
error
in
na
7.58
- 5.52
- 5.85
- 8.49
1.74
4.24
6.94
- 1.16
- 3.30
5.89
- 3.83
Maximum
% error in
expansion"
13.7
- 15.0
- 26.0
- 24.8
4.7
5.0
1.9
- 1.6
- 3.8
4.1
- 2.7
 Compared  to values  in Table 44.

 b[(Observed depth -  predicted depth)  100/observed depth].  Maximum
  observed  error  in predicted expansion up to a total bed expansion
  of  70% above  fixed  bed  depth.
models appropriate to media of about the same sphericity as that used
in their development.  Thus, they should be used with caution on
media from other sources with potentially different sphericity.  Future
work should include development of simple direct measures of sphericity
and collection of additional data on the effect of sphericity on the
drag coefficient and on the expansion models.

Since the prediction of expansion of the uniform sands was considered
acceptable, the models were used to predict the expansion of the
three graded (A, A-2, and C) sands previously described in Fig. 60
                                  296

-------
and for graded sand observations reported by Amlrtharajah [4].  The
results of the prediction are summarized in Table 47.  The procedure
consisted of the following steps:  (1) calculation of the average
grain diameter by the inverse definition [Eq. (38)], (2) calculation
of the Galileo number from the properties of the media and fluid,
(3) calculation of Re from Eq. (73) and V± from Re^ (4) calculation
of n from Eq. (71), (5) calculation of e at each superficial flow
rate, V from Eq. (28) using V. and n calculated above, and (6) cal-
culation of expanded bed depth from Eq. (2).

It is evident from Table 47 that the predicted bed depths are all
higher than observed, and the prediction is not too good.  An attempt
was made to determine the reasons for the consistent over-prediction.
One cause is the choice of e0 used in Eq. (2).  Values of eo of 0.42
were used for the graded sands of Boss and 0.41 for those of
Amirtharajah.  These choices were based on porosities by the
graduate cylinder technique for graded sands described previously
and values reported by two investigators.  If a value of 0.44 had
been selected as determined by the column technique (Table 37),
the prediction would have improved.  The arithmetic mean diameter
[Eq. (39)] consistently yields a larger diameter than the inverse
diameter [Eq. (38)], as evidenced by Table 35.  If the larger
diameter defined by Eq. (39) had been used, the prediction would
be improved.  A spot check of the effect of these two factors on
prediction for three graded sands in Table 47 indicates that the maxi-
mum percent errors reported would be reduced about 5% (e.g., from
-15 to -10%).

From this analysis, it is evident that the prediction adequacy is
sensitive to choice of eo and mean diameter.  Acceptable predictions
seem possible if the column technique is used to evaluate e  and the
arithmetic mean definition [Eq.  (39)1 is used to calculate mean grain
diameter.

The expansion of the graded sands was also calculated using the
incremental approach presented in some engineering textbooks[46].
This approach calculates the expansion of increments of the media
between adjacent sieves and sums the expanded depth of the increments
to determine the total expanded  depth.  The results of the incremental
approach were no better than, and in  some cases worse than, those
presented in Table 47.

In view of the relative inaccuracy of prediction of expansion of the
uniform coals, no  attempt was made at this  time to  test the prediction
accuracy of  the models  for graded coals.

Minimum Fluidization Velocity of all Media

The minimum  fluldization velocity  (Vmf)  can be defined  in a number of
ways.  For uniform media which  fluidizes  sharply at  a particular flow
                                  297

-------
 Table 47.  Prediction of expanded bed depths  for graded sands using
           models developed  for uniform sands and average diameter
           based on  the inverse definition  [Eq. (38)].
Sand
From
A
A-2
C
From
A
A
A
A
A
B
B
B
B
C
C
Avg
dia,
nun
Boss [14]
0.686
0.716
0.947
Water
temp»
°C

17
26
22
Ga

4,383
7,736
14,838
R6i

56
76.9
110.4
Vi,
fps

0.2926
0.3092
0.3353
n

3.640
3.424
3.191
Maximum
% error
in expanded
depth3

- 7.10
- 6.6
- 17.3
Amirtharajah [4]
0.612
0.612
0.612
0.612
0.612
0.919
0.919
0.919
0.919
0.688
0.688
15
13.5
30
30
30
15
15
30
30
22
35.5
2,824
2,613
5,773
5,773
5,773
9,599
9,599
19,623
19,623
5,725
10,401
43.9
42.0
65.3
65.3
65.3
86.6
86.6
129
129
65.0
90.6
0.270
0.269
0.281
0.281
0.281
0.354
0.354
0.370
0.370
0.298
0.309
3.818
3.850
3.534
3.534
3.534
3.347
3.347
3.096
3.096
3.537
3.316
- 9.5
- 11.5
- 14.9
- 16.2
- 16.2
- 15.8
- 17.8
- 15.8
- 16.8
- 17.4
- 17.7
a[(Observed depth - predicted depth) 100/observed depth], over the
 full range investigated, generally to an expanded depth of about
 50% of fixed bed depth.
                                 298

-------
rate, it is frequently defined as the point of intersection obtained
through extrapolation of the two linear sections of the envelope curve
of head loss vs superficial velocity such as Figs 73 and 74.  The
values of Vmf based on this definition for garnet sands have been pre-
sented in Table 40.  However, this definition is not meaningful for
graded media because the coarser sizes of media comprising the bed
are not fluidized at the Vmf defined in the above manner.

It is also difficult to define Vmf on the basis of first visual ap-
pearance of complete fluidization because it is subject to the ob-
server's visual definition of complete fluidization.  In view of these
difficulties and the fact that all media studied were graded in size
to some extent (even the uniform media), the minimum fluidization
velocity was defined as that flow rate required to achieve 10% bed
expansion.  This expansion is close to the minimum rate at which the
bed first appears fluidized.  The Vmf values thus determined for the
three media studied are presented in Fig. 88.  These data are based
on the expansion studies for single uniform media only in Table 33
(Series C-l and 3 and D-l) and in Table 31 (Series A-13, 14, and 15).

This type of data is believed to be useful in determining minimum
backwash rates to ensure fluidization of all media comprising the bed.
If the sieve analysis of each media is available or specified, the
backwash rate needed to achieve fluidization of the coarse  sizes in
each media should be provided.  Furthermore, it would  appear  that
media should be specified so that the coarse sizes of  each  media
comprising the bed are fluidized at roughly the same minimum
fluidization velocity.  Thus, the entire bed will become  fluidized
simultaneously.  The total bed expansion can then be determined by
summing up the calculated expansions of the individual media.

Unfortunately, the experiments depicted in Fig. 88 were  not conducted
over a broad enough range of temperatures to present similar  empirical
data for other temperatures.  The effect of  temperature  on  Vmf  can be
judged from various models  for V £ such as Eqs.  (51) through  (54).

Data were collected on graded sand A and graded coal A at four dif-
ferent temperatures from  16  to 30 °C.  To illustrate the effect of
temperature on V^, the values of V  ^  to  achieve  10% expansion  of
these two media were determined.  The  results  are plotted in  Fig. 89.
It is evident  that  temperature does have  a  distinct effect  on Vmf
as would be expected in the transitional  or laminar fluid regimes.
These curves could be used  to select a rough temperature correction
factor to be applied to the data  of Fig.  88  in selecting Vmf  for
other temperatures.  Inspection of Fig. 89  shows  that  change  in Vmf
is not a  linear inverse function  of viscosity  as would be expected
in the transitional fluid regime.  Only in  the laminar regime would
the  relation be  linear.
                                  299

-------
                                     COAL p = 1.7          o
                                     SILICA SAND p = 2.65   D
                                     GARNET SAND p« 4.13
      .0  0.25 0.5  0.75   1         1.5        2
                           MEAN SIEVE SIZE, mm
       2.5
 Fig.  88.  Minimum fluidization velocity, V  f, to achieve 10% bed
          expansion at 25 °C.            m
0.04
0.03
0.02
0.01 h
                  10
      20
TEMP,  °C
30
          1.0 1
          0.5
                                                              §
                                                              u
                                                              i/>
                                                              O
                                                              u
                                                              i/>
                                                              £
Fig. 89»  Effect  of  temperature on V - for sand and coal and en
          -• - -*    viscosity of water.
                               300

-------
Comparison of Fig. 89 with Table 40 shows reasonable agreement with
change in Vmf with temperature.  For example, the average ratio of
Vmf values for garnet at 25 °C and 17 °C is 1.14 in Table 40.  The
same ratio for sand and coal in Fig. 89 is 1.17.

It should be noted that the coarse sizes in a given media control
the point of complete minimum fluidization, and skimming the fines
will not alter the V c.
                    mi

                      Intermixing Observations

Garnet and Silica Sand

The objective of the Intermixing experiments was to test the validity
of the bulk density approach for the prediction of intermixing of
the small-sized dense garnet sand and larger-sized less dense silica
sand.  A uniform-sized garnet sand and various uniform-sized silica
sands were first fluidized individually, and then the two-component
mixtures of the silica sands and garnet sands were fluidized to ob-
serve their intermixing behavior.

The data collected for the single-component fluidization experiments
included flow rate and bed height.  From the individual component
fluidization data the following values were calculated:

1.  Flow meter readings were corrected to give corrected flow rate
    using the appropriate calibration equation.

2.  From the bed height reading, the average porosity of the bed
    at various flow rates was calculated from the known weight of
    media in the column, the column cross-sectional area, and the
    particle density.

3.  The average bulk density was then calculated at the same flow
    rates from the above porosity values and the solid and  fluid
    density by Eq. (41).

The computed values of flow rate and bulk density  from single-component
fluidization data for each  media studied are shown in Fig.  90.  An
interesting and important fact  can  be observed  on  Fig. 90.  The slope
of the garnet sand curve is steeper than any of the silica  sand curves.
Thus, maximum bulk density  differences  (garnet  — silica sand) occur at
the lowest flow rates.

The data collected for the  two-component  fluidization experiments
were  the bed height, flowmeter reading,  the visual observations  of
relative media concentrations  at various depth  in the bed.   The
results of these  two-component fluidization experiments  are shown in
Figs. 91 through  94.
                                  301

-------
                                                          - 2.4
                            o GARNET SAND-50+ 60
                            D SILICA SAND-20+ 25
                              SILICA SAND-30+ 35
                            • SILICA SAND -35 + 40
                            A SILICA SAND-40+ 45
                                                          -  1.0
                10      20     30      40

                      FLOW RATE, V, gpm/sq
Fig. 90.  Bulk density vs flow rate for garnet sand and silica sand.
Before the intermixing observations were made,  the two-component mix-
ture was fluidized and contracted to a fixed-bed state very slowly.
The data presented in Figs. 91 through 94 were  then collected  during
the expansion of the two-component mixtures.  After each incremental
increase in flow, sufficient time was allowed to reach equilibrium
conditions before the intermixing observations  were recorded.  The
expansion was carried up to about 200%.   The  major problem encountered
in collecting and presenting this type of data  was that it was difficult
                                  302

-------
OJ
o
         60
         50
         40
       g30
       x
       Q
GARNET SAND S
 SILICA SAND 0
 INTERMIXING D
         120
         10
            0      10     20      30     40 NO FLOW
                   FLOW RATE, V , gpm/sq ft

        Fig. 91.   Intermixing of -50+60 garnet sand
                  and -2OI-25 silica  sand.
60
                                             50
                                           c 40
                                           **
                                          x
                                          § 30
                                          x
                                          o
                                          LU
                                          CO
                                             20
                                             10
GARNET SAND S
SILICA SAND   0
INTERMIXING  D
                                             Fig.  92.
          10      20      30      40 NO FLOW
          FLOW RATE, V , gpm/sq ft

          Intermixing of -50+60 garnet  sand
          and -30+35 silica sand.

-------
U)
                 GARNET SAND C3
                  SILICA SAND CZ3
                 INTERMIXING O
          Fig. 93.
10      20     30     40 NO FLOW

  FLOW RATE, V , gpm/sq ft

Intermixing of  -50+60 garnet sand
and -35+40 silica sand.
60
                                                             50
                                                             40
                                                          O30
                                                          i
                                                          Q
                                                             20
                                                             10
                                             GARNET SAND  E3
                                            —  SILICA SAND  E3
                                              INTERMIXING  I—I
                                                                      10     20     30     40 NO FLOW
                                                                     FLOW RATE, V , gpm/sq ft
Fig. 94.   Intermixing of -50+60 garnet sand
          and  -40+45 silica sand.

-------
at times to decide in which regions of the expanded bed one component
was the sole component and in which the components were intermixed.
Three regions usually existed — one region at the top  and one at the
bottom of the bed where the components were clearly or closely the
sole component of the layer, and one between these two separate layers
which was a region of intermixing.  Within this region of intermixing,
the concentration of each component decreased with distance away from
the adjacent region where it was the major component.  The interfaces
between the three different regions are distinguished on the figures
by a solid line which indicates a sharp interface or a dashed line
representing a diffuse interface over a bed depth of 2 to 4 in.

After the two components were expanded to the maximum expanded height,
the flow of water was quickly shut off, and the bed was also allowed
to settle.  The stratification of the bed after this settling is also
included in Figs. 91 through 94.

The analysis of Figs. 91 through 94 indicates a trend for the garnet
sand to occupy the lower layer of the bed at low flow rates, then as
the flow rate increases, the two-component mixtures  are intermixed.
At still higher flow rates, the garnet sand occupies the upper portion
of the bed, and the silica sand occupies the bottom  of the bed.  These
trends would be expected from the bulk density plots of the single
component data as shown in Fig. 90, because the bulk density of the
garnet sand decreased more rapidly than the silica sand as the flow
rate increased.

An exception to this trend was noticed in the fluidization of the
-20+25 silica sand and -50+60 garnet sand mixture where the silica
sand occupied the lower portion of the bed at all flow rates.  This
can be explained by considering the minimum fluidization velocities
of the silica sand and the garnet sand.  The minimum fluidization for
the -50+60 garnet sand was observed  to be 3.0 to 3.5  gpm/sq ft  (0.007
to 0.008 fps).  The silica sand minimum fluidization velocity was
approximately 7.5 gpm/sq ft  (0.017 fps) for the  -20+25 size.  There-
fore, when the fluidized mixture of  -20+25 silica sand and garnet
sand was contracted slowly,  the -20+25 silica sand reached the fixed
bed state while the garnet sand was  still fluidized  and before the
bulk density of the garnet sand was  sufficiently greater  than the
silica sand to occupy the bottom layer of the bed.   Just below mini-
mum fluidization for the -20+25 silica sand, it was  noticed that the
garnet sand, which was still fluidized, displaced a  small portion of
the silica sand that was in  a fixed  state up into the intermixed
layer.

The garnet sand was below the silica sand at all flow rates  for  the
-40+45 silica sand and -50+60 garnet sand two-component mixture.
The maximum flow rate of this two-component mixture  was 40  gpm/sq  ft
 (0.089 fps).  At this flow rate,  the garnet s-and still had  a higher
bulk density than the -40+45 silica  sand  (from  single-component  data,
                                  305

-------
 Thus, as predicted by the bulk density approach,  the garnet sand
 should occupy the lower portion of the bed.

 The sensitivity of results was hampered by the following factors.

 1.  The uniform sands used were not actually unisized and of con-
     sistent shape.  Because of this, there would  be a tendency for
     stratification within each individual media,  and a bulk density
     gradient would exist in each individual  media.   This was sup-
     ported by the observation for all of the uniform media when ex-
     panded and then contracted slowly, or when expanded and allowed
     to settle after the fluid flow was stopped.   There was a slight
     but noticeable difference in particle size between the top and
     bottom layers.

 2.  Because of the physical properties of the fluid and solids, the
     fluidization of the components occurred  in the  transitional
     regime of flow, and mixing and circulation patterns existed in
     the bed during fluidization.  This would tend to diminish the
     bulk density gradients within the individual  and two-component
     mixtures.

 3.  The distribution of flow into the fluidizing  column from the
     calming section was not perfectly uniform.  However, it was
     quite good with short circuiting of upward flow usually limited
     to 2 to 3  in.  above the entrance and rarely extending 6 in. up
     the column.

 There are three  major conditions of interest in the relative location
 of the garnet  sand component:

 1.   stratification of the garnet sand component in  the bottom of the
     bed  with or  without intermixing of garnet sand  and silica sand
     in layers  above.

 2.   maximum intermixing of  the  garnet sand with silica sand,  and

 3.   stratification of garnet  sand  in the  top layer  with or without
     intermixing  in layers below.

The  following  bulk density  differences, Table 48, were obtained from
Fig.  90  and  the  appropriate two-component intermixing  figures  (Figs.
91 through  94) for the  three  conditions.  The ratio of the  diameters
of garnet sand to  silica  sand is also given  below.  From Table  48
or from Figs.  90 and  91  through  94, maximum  intermixing of  the  two
components  (Condition 2)  does not  occur at zero bulk density difference
as would be expected, but at  a slightly positive bulk  density differ-
ence,  about 3  to 8  Ib/cu  ft.  These  observations are in agreement
with  the experimental results of Le Clair [74] who  found that  the
velocity for a homogeneous mixture  as predicted from single-component
data was greater than the observed velocity where homogeneous mixing
occurred.
                                  306

-------
Table 48.  Bulk density difference, Ib/cu ft (garnet-silica sand).
Silica sand
   media
Condition
   (1)
Condition
   (2)
Condition
   (3)
,   Silica
m  sand
d_ Garnet
m     ,
    sand
  -20+25
 Did not
 occur
  8 to 3
  < 12
  2.84
-30+35
-35+40
-40+45

> 20
> 10
> 5

15 to 5
8 to 4
10 to =" 0

< 9
< 6
Did not
occur
2.04
1.70
1.47

It is evident from Table 48 that a single value of bulk density
difference cannot be readily selected which could be used by the
design engineer to ensure the desired degree of stratification or
intermixing between garnet sand and silica sand.  For  example, the
bulk density differences (garnet-silica sand), which ensure that
Condition 1 or 3 will exist, decreased with lower diameter ratio
(silica sand/garnet sand).  The lower bulk density differences, indi-
cated in the table for Condition 1 and 3, occur at higher flow rates
and higher bed porosities.

Based on the data  in Table 48,  if  one wanted  to  ensure that  some
garnet  sand would  always  remain on the bottom regardless  of  the back-
wash rate of the filter bed,  one would need  to select  the media  so
that the ratio of  the bottom silica sand  size to the bottom  garnet  sand
size is not more than 1.47.   The next  larger ratio could  be  used  (1.70)
providing that the backwash  rate was  limited to  15 to  20  gpm/sq  ft
(0.033  to 0.045 fps) or  the  operator would need  to allow  for a slow
contraction of the fluidized bed  after backvashing to  achieve restrati-
fication.

Attention should be  drawn to the  converging  nature of  the garnet  sand
and  silica  sand curves  in Fig.  90.  Because  of this converging nature,
excessive backwash rates  result in increased tendencies to  inter-
mixing  and bed inversion.  This fact  should  be considered in selecting
the media and backwash  rate  for dual-  and multi-media filters.

In view of  the apparent inadequacies,  or insensitivity of the equal
bulk density  approach  to  the prediction of intermixing, an attempt was
made  to utilize  the  intermixing theory proposed by Camp et al. [26].
                                   307

-------
 This theory as stated in Eq.  (49)  suggests that mixing will occur if
 the bulk density of the lower bed  of smaller, more dense garnet grains
 is less than the density of the larger,  less dense upper particles
 minus a drag force term for the upper grains.  Figure 95 shows the
 results of the calculations.

 Figure 95-A shows that the -204-25  mesh silica sand should intermix
 at all flow rates since its particle density less the drag term is
 still greater than the bulk density of the garnet,  top or bottom
 layer.   This intermixing was  evident in  Fig.  91.

 Similarly,  Fig.  95-B for -40+45 silica sand would predict complete
 intermixing if the sand grains  were spherical.  However,  if they were
 cubical in  shape,  the drag term would be increased,  and intermixing
 would be only partial.   The sand grains  used in this study are rounded
 in shape and would not be cubical  in sphericity.   Thus, intermixing
 should be complete.   However, Fig.  94 shows that  most of the  garnet
 sand remained on the bottom of  the  bed at all flow rates for  this
 sand and garnet  combination.

 These observations do not prove or  disprove the validity of the inter-
 mixing  model of  Camp et al. [26].   An adequate  measure of sphericity
 would be needed  to test the model.   The  unavailability of such a
 measure is  the same  weakness preventing  good  prediction of bed expan-
 sion for nonspherical particles discussed previously.   Until  this
 weakness is resolved,  the intermixing model of  Camp  et al.  [26]  is
 no more useful than the equal bulk  density model  of  Le Clair  [74].

 Silica  Sand and  Coal

 Bulk density data  and  intermixing observations  for  the silica sand
 and coal were collected in the  same manner previously described for
 garnet  and  silica  sand.

 A  full  range of  uniform media was tested in a series of tests  using
 the  2-in. fluidization  column.   The uniform sizes tested  ranged from
 -10+12 mesh to -40+45 mesh for  sand and  from  -4+7 mesh to -25+30  mesh
 for  coal.   A known dry weight of a  particular uniform filter medium
was  placed  in the  column.  The  expanded  bed height vs  flow rate was
observed.   Porosity  at  any flow rate was  calculated  from  the known
weight of the medium, particle  density,  and expanded bed  height.   Bulk
density  for each flow rate was  calculated  by Eq.  (41)  and then plotted
 in Fig.  96.   Bulk  density,  as defined  in this study,  is actually  the
average  fluid  and  particle composite density value within a filter
bed  cross section.  At  zero flow rate, the  sand bulk density was
approximately  1.9  g/cc while the coal bulk density was  approximately
1.3  g/cc.

Figure 96 illustrates that, with an increasing  flow  rate,  sand  bulk
densities decrease much more rapidly than  the coal bulk densities.
An extreme bulk density decrease of  from 1.90 g/cc at  zero  flow rate
                                  308

-------
to

UJ
O
 u
i
UJ
      2.5
      2.0
1.5
      1.0
      0.5
      2.5
2.0
       1.5
       1.0
       0.5
                                         p  SILICA SAND
                                    DRAG FORCE ON  SAND
                                    ASSUMING SPHERICAL
                                     1HAPE
                                 DRAG FORCE ON SAND -
                                 ASSUMING CUBICAL SHAPE
GARNET - BOTTOM OF BED
«.
pb
GARNET - TOP OF BED
              [A)  -20 + 25 SILICA SAND AND -50 + 60 GARNET SAND
                                                        AT 25 °C
                                     p$ SILICA SAND
                                          DRAG FORCE ON SAND
                                           ASSUMING SPHERICAL
                                                  SHAPE
                            DRAG FORCE ON SAND -
                            ASSUMING CUBICAL SHAPE
                                     P.  GARNET - BOTTOM
                                        OF BED
                                              GARNET - TOP OF BED
              (I)  -40 + 45 SILICA SAND AND -50+60 GARNET SAND
              ^—-                                            _ ft _
                                                         AT 25 "C
         0.00    0.02     0.04     0.06     0.08     0.10

                          SUPERFICIAL VELOCITY, fps
  Fig. 95.  Intermixing of silica sand and coal according  to the
            model of Camp et al. [26].
                                 309

-------
      -25+30
      -30+35
      -35+40
      -40+45
                                             F-7+8

                                             F-8+10
                                             F-10+12
 65 —
                                                                u
                                                                u
                                                                o>
                                                      l/l
                                                      z
                                                      LU
                                                      Q
Fig. 96.
            20       40      60       80
                 FLOW RATE, V ,  gpm/sq ft
Bulk density vs flow rate for coal and  silica  sands
(data points not shown on all curves  for  drafting  con-
venience) .
                              310

-------
to 1.15 g/cc at 50 gpm/sq ft was exhibited by the -40+45 mesh sand.
The coal bulk densities showed less dependence on flow rate,  especially
coarser coals down to 12 mesh.  The general pattern is similar to that
for garnet and silica sand presented previously in Fig. 90.

A bulk density comparison between uniform media of the same size
range but different origin  (i.e., -10+12 mesh coal A and -10+12
mesh coal F) is also plotted in Fig. 96.  The bulk density difference
of these identically-sized uniform coals was attributed to dif-
ferences in specific gravity and particle shape of the two source
samples.  Limited data restricted such comparisons of uniform
media from different sources to the two uniform media above.

A large bulk density difference also exists between the majority
of the sand and the majority of the coal media*  If the back-
washing flow rate is limited to 30 gpm/sq ft, only the fine -30
+45 mesh sand approaches the bulk density of the coarse coal.

Intermixing observations were limited to -4+7 coal F and the  various
uniform sands shown in Fig. 96.  Again, it was desired to test the
validity and sensitivity of the equal bulk density theory of inter-
mixing of Le Glair [74].  In retrospect, it would have been better
to use several narrow, size ranges for the coal.  Use of only the
broad -4+7 mesh coal in the intermixing observations led to incon-
clusive results regarding the bulk density difference associated
with partial or complete intermixing.  Additional work in this area
should be conducted.

Expansion of Graded Dual-Media Filters

This section presents  some  general characteristics  associated with
the expansion of dual-media filters.

(1) An example of the  expansion plot for  a graded dual-media filter
bed is found in Fig. 97.  The bed height  ordinate  shows  that approxi-
mately equal 12-in. quantities of  sand A  and  coal A were systematically
expanded.  The similarity  shown between the  expansion  rate of sand A
and coal A resulted, because  the coal  average particle  size was  such
that the V  , values were almost the  same  for  sand A and  coal A.
          ml
Figure 97 illustrates  that the predicted  expansion,  obtained by
adding bed expansion of  the individual  components,  closely approxi-
mated the experimentally determined  expansion of  the dual-media
filter.  The results of Fig.  97 support the work  of Le  Clair  [74],
who, in restating the  cell theory,  suggested  that  in all cases each
fluidized particle within  a medium has  a  definite,  surrounding cell
of fluid  at  a particular flow rate.  Introducing particles of dif-
ferent densities does  not  influence  the size  of  the fluid cells  for
the particular flow rate.   Therefore,  the expansion of the individual
components can be summed to predict  dual-media  filter  expansion.
                                  311

-------
   42
   36
   30
   24
i
o
a  is
   12
              SUM OF TWO
              SINGLE MEDIA
                                     O SAND A
                                     O COAL A
                                     A DUAL MEDIA AA
                     10       15      20       25

                     FLOW RATE, V, gpm/sq ft
30
 Fig.  97.  Expansion vs flow rate of dual media AA and  the two-
          component media at 22 °C.
                              312

-------
(2) The expansion of dual media A-2^2 an<* fc^e tw£>-component media are
presented in Fig. 98.  Because sand A2 became fluidized first, ex-
pansion of the dual-media filter at lower flow rates was close to
the amount of sand expansion.  At higher flow rates, above fluidi-
zation for coal C2, expansion of dual media A2C2 was close to the
sum of the expanded depths of the two single media.

It is apparent from Fig. 98 that sand A£ begins to expand at  about
7 to 8 gpm/sq ft and reaches 107. expansion at about 12 gpm/sq ft.
Coal C2 on the other hand begins expanding at about 15 to 16  gpm/sq ft
and reaches 10% expansion at about 22 gpm/sq ft.

The most important fact to be concluded from Fig. 98 is that  if the
sand and coal of a dual-media filter are not selected to fluidize
at about the same flow rate, one media may remain fixed (or nearly
fixed) while the other is fluidized.  The fixed bed portion may not
be cleaned adequately during the backwashing.  For example, if dual
media h-2^2 were expanded 20%, the minimum normally required for
adequate backwashing, the coal would be expanded only about 5%.  At
5% expansion, the coal is essentially in a fixed-bed condition and
would not clean well.  Furthermore, to achieve 20% expansion  of the
coal in this dual media would require a flow rate of 27 gpm/sq ft,
and the total bed expansion would be about 35%.  Thus, observation
of a total bed expansion of 20% for this dual media would not
necessarily mean that both media would receive an adequate backwash.
                                  313

-------
    35
   30
   25
                SUM OF TWO
                SINGLE MEDIA
                                       O SAND A2
                                         COAL C2
                                       a DUAL MIDIA
                10         20         30
                       FLOW RATE, V, gpm/sq ft
40
Fig.  98.  Expansion vs flow rate of dual media &2C2 and the two~
         component media at 22 °C.
                             314

-------
                  XI.  EFFECT OF MEDIA INTERMIXING ON

                         DUAL-MEDIA FILTRATION


                             Introduction

Dual-media filters composed of anthracite coal over silica sand are
widely used in water and wastewater filtration because they achieve
greater water production per filter run than sand filters if other
filtration conditions are the same.  Furthermore, because of greater
potential production per filter cycle, the percentage of filtered
water used in backwashing is less for dual-media filters than for
sand filters.

The coal and sand of a dual media filter will sometimes intermix at
the interface, depending on the size, shape, and density of the two
media at the interface, the rate of backwash, and the backwash valve
closure rate.  The most important parameter affecting the amount of
intermixing is the relative size ratio of the sand and coal at the
interface.  The media may be so graded as to maintain a sharp inter-
face (coal size to sand size ratio at the interface of about 2 to 1)
or to allow a substantial zone of intermixing (coal size to sand size
of about 4 to 1).

There are two conflicting ideas concerning the desirable amount of
intermixing.  Some researchers feel there should be no mixing at the
interface, with the sand acting as a polishing filter after the coal
roughing filter.  Others feel that intermixing results in a more
uniform decrease in grain size with depth and allows more efficient
use of the storage space in the media and, thus, longer filter runs.
An intermixed bed is a closer approximation of the ideal coarse to
fine filter bed and thereby eliminates an impervious  layer that might
build up at a sharp interface.

                  Objectives and Scope of This Study

Because of the differences of opinion concerning the  desirability of
intermixing in a dual-media filter,  this study was undertaken to show
what effect, if any, the intermixing has on the performance of dual-
media filters.  Performance differences, if they exist, will be shown
by measurements of head  loss development and  effluent quality versus
time during filtration.

It must be kept in mind  that this  study was not meant to be a com-
parison of dual-media  and single-media filtration nor a study to
determine the optimum  amount of  intermixing;  rather,  it is meant to
show how the media  intermixing or  non-intermixing in  dual-media fil-
tration affects performance.
                                  315

-------
                       Experimental Investigation

 Apparatus and General Approach

 The system used in this study consisted of three, 4-in.  inside diam-
 eter plexiglass filter columns as shown in Fig.  99.   In  one column,
 the coal and sand were placed together and allowed  to mix  as  they
 might.  In two other columns, operated in series, identical coal and
 sand were placed separately.   Two gradations  of  coal  and sand media
 were used, one which resulted in a mixed interface, dual-media bed
 and one which resulted in a sharp interface dual media.

 Prior to further detailed description of the  experimental  investiga-
 tion some terms must be defined for clarity.  The term "mixed inter-
 face media" refers to the media which exhibited  intermixing of the
 anthracite and sand at the interface when both media  were  placed in
 column 1.  The term "sharp interface media" refers  to the  media which
 exhibited no intermixing of the anthracite and sand at the interface
 when both were placed in column 1.   The term  "combined media" refers
 to the sand and anthracite media when placed  together in column 1.
 The term "separate media" refers to the same  amount and  gradation of
 sand and anthracite placed in separate columns.  The  same  anthracite
 as in column 1 was placed by  itself in column 2.  The same sand as
 in column 1 was placed by itself in column 3.

 A filter run consisted of measuring the head  loss buildup  and efflu-
 ent quality for the combined  media  and the head  loss  buildup  and
 effluent quality for the separate media when  the separate  and combined
 media filters were operated concurrently.   Comparisons between the
 results  obtained with  mixed and with sharp interface  media in subse-
 quent filter runs were then made to evaluate  the effects of inter-
 mixing on filter performance.

 The filters  were operated as  pressure  type filters with  the pump
 applying approximately 25 psig  pressure  to the top of filter column
 1  and column 2  (Fig. 99).   The  suspensions  of particulates were
 pumped from  the supply to columns 1 and  2.  From the  separate coal
 (column  2),  the water  flowed  to column 3, containing  the separate
 sand.  The filtered water coming from  column  1 and column  3 flowed
 through  a pressure regulator, a needle valve, a rotameter,  and to
 waste.

 Constant  rate filtration  was  achieved  by use of Fisher,  type 95L
 pressure  regulators  (Fisher controls Company, Marshalltown, Iowa).
 These  regulators were used  to reduce any incoming pressure to a con-
 stant  exit pressure of 4 psig.  This constant 4 psig pressure leaving
 the regulator was applied  to a fixed effluent needle valve to achieve
 a constant flow.

This  system provided a reasonably constant flow,  although slight
variations in flow were observable on the effluent rotameter which
                                 316

-------
     PUMP
FILTER INFLUENT
BACKWASH
EFFLUENT
                         _BACKWASH_
                          EFFLUENT
                     I	
             NEEDLE VALVE
                 COLUMN
                   NO. 2
                                             BACKWASH
COLUMN]
 NO. 3
              BACKWASH LINE
   NEEDLE VALVE
                    Fig. 99.  Schematic diagram of apparatus.

-------
 had a full-scale capacity of 0.78  gpm.   The rotameter  discharged  to
 waste.

 Tap water at normal main pressure  was used  for  backwash.  The back-
 wash system is shown in the plumbing diagram, Fig.  99.

 Head loss buildup was measured using a multiple tube manometer con-
 taining four single leg manometers (Model 33KB35, Multiple Tube Ma-
 nometer,  Merian Instrument Division, The Scott  and  Fetzer Company,
 Cleveland,  Ohio).  A 1/4-in.  OD copper tube was connected from the
 top of a  filter housing to the bottom end of one manometer, while
 another 1/4-in. OD copper tube was connected from the  bottom of a
 filter housing to the top end of the same manometer.   The manometer
 operates  on the principle of  a U-tube with  one  leg  of  water and one
 leg of mercury.

 Head loss buildup was observed by  the differences in the readings of
 the mercury level at successive time intervals  during  a filter run.
 Initial head loss through each filter could not be  readily measured,
 but head  loss increase with time was readily and precisely observed.

 Before and  after every run,  the filters  were backwashed.  The filters
 were expanded to 50% during the backwash and allowed to wash thor-
 oughly.   When a cake of solids formed on the surface of the media,
 air scour was sometimes used.   After a thorough backwash, the back-
 wash valve  was closed rapidly.   Then the bed was shocked momentarily
 by  rapidly  opening and closing the  backwash valve several times in
 succession.   This procedure was used to  insure  an equally dense bed
 for every filter run.   The  bed depth was recorded after every wash
 for each  run and did not  vary.

 Because the patterns  of head  loss  buildup and effluent quality were
 of  more importance than absolute quality, the filters were operated
 at  7  gpm/sq ft.   This  high  filtration rate was  also selected to en-
 courage deep penetration  of some solids  into the bed, hopefully
 through the  interfacial region to  allow  the  effects of the interface
 to  be observed.

 Filter Media

 Two different  dual media  were  used  for this  study.  One dual media
 was  selected in order  that  intermixing would occur at the interface.
 The second dual media was selected  in order  to  produce a sharp, well
 defined interface.  The sharp  interfaced dual media was intended to
 act  as a control,  to  show what  differences  in head  loss and filtrate
 quality could be  attributed to  the experimental apparatus and opera-
 tion.  However, the sharp interface media was not a control in the
 strict sense of the word, because the sharp  interface media and the
mixed interface media were not  operated at  the  same time.
                                 318

-------
Any effects on head loss development and effluent quality produced  by
the method of operation or the equipment (presenting two surfaces to
the flow, flowing through an underdrain and plumbing,  etc.)  had to
be determined.  The filtration with the sharp interface media was
done to illustrate these effects.  As there was no mixing of the
media in column 1 (Fig. 99), there would be no effect of intermixing
on the performance of the filter in column 1.  There was obviously
no intermixing between the coal in column 2 and the sand in column  3
and no effect due to intermixing.  A comparison of the results of the
combined media filtration with the separate media filtration using
the sharp interface media would show what effect the operating proce-
dure and equipment had on the results.  This observation would then
be useful in interpreting the data collected using the mixed inter-
face media in subsequent filtration runs.

The dual media, selected to have a well intermixed interface, was
similar to a media that is commonly specified for water treatment
plant filters.

The coal was "Philterkol" from the Reading Anthracite Coal Company,
Pottsville, Pennsylvania.  The sand was from the Northern Gravel Com-
pany, Muscatine, Iowa.  Fines from the coal and sand were skimmed
from the media after hydraulic gradation of the media by backwashing.
This was done because  it was felt these fines might cause problems
and because skimming is commonly practiced in  filter plant construc-
tion.

The detailed  skimming  procedure was as  follows.   Approximately  30 in.
of  sand  or coal was placed  in the plexiglass column.  The media was
fluidized  to  approximately  50% expansion and allowed to stabilize.
The backwash valve was closed rapidly,  and  the media was  allowed to
subside.  The  top inch or so was  skimmed off by  a siphon.  The  pro-
cedure was repeated until 10  to  20% of  the  finer  media  was removed.
The media was  then removed  from  the column  and dried in an oven.  A
sieve analysis was run on the  skimmed coal  and sand.  The effective
sizes and  uniformity coefficients  of  the  skimmed media  were  as
follows;

      Sharp interface media
           sand   ES =  0.85  mm
                  UC -  1.29
           coal   ES =  0.91  mm
                  UC =  1.45
      Mixed interface media
           sand   ES  =  0.46  mm
                  UC =  1.49
           coal   ES  =  0.92  mm
                  UC  =  1.60

 The size ratio at the  interface is the most important factor deter-
 mining the amount of  intermixing of coal and sand.  In this study,
                                  319

-------
 the 99% finer size by weight (dgg^)  was  used  to approximate  the
 coarsest size coal, and the d^o size was used to approximate the
 finest sand.   The size ratio at the  mixed interface was 5.34.  The
 size ratio at the sharp interface was 2.32.   The size ratios based
 on the dgo% for the coal and the dio% for the sand were 4.05 for the
 mixed interface media and 1.93  for the sharp  interface media.

 The specific  gravity of "Philterkol" ranged from 1.65 to  1.7 on
 various shipments whereas the sand has consistently been  2.65.

 Source of Suspensions

 Several different suspensions of participates commonly encountered in
 filtration were filtered in an  attempt to ascertain what  effect, if
 any,  intermixing of the media in dual-media filtration has on filter
 performance.   It was felt that  different  types of suspensions might
 have different transport and attachment mechanisms for the removal
 of solids,  and the effect of intermixing  for  various typical filtra-
 tion situations might be different.

 Five different suspensions were filtered.  They were an iron floe,
 lime-soda ash softening precipitate,  aluminum sulfate coagulated and
 settled trickling filter effluent, activated  sludge settled  effluent,
 and trickling filter settled effluent.

 Iron floe.  Ferrous sulfate  was mixed with Iowa State University tap
 water to prepare an influent suspension containing precipitated iron
 floe  for the  filtration study.   A stock feed  solution of  0.2M  fer-
 rous  sulfate  in an acid solution  of  approximately 0.1 N HCL was made
 up in sufficient quantity to fill a  20-liter  feed bottle.

 To achieve  a  mixing tank effluent of 9 to 9.5 mg/1 of iron.  17.4
 ml/min of the stock ferrous  sulfate  solution was fed into 5.76 gpm
 of tap water.   The iron solution  was dripped  from a constant head
 capillary feeder into  a mixing  tank  to achieve  the desired iron con-
 centration.   The iron  solution was mixed by a  paddle mixer in the
 reaction tank.

 The type of precipitate  formed by the addition  of ferrous sulfate to
water depends  on the pH,  alkalinity,  and temperature of the water
 and upon the  time  allowed for reaction.  It may consist of Fe(OH)o,
FeC03,  or Fe(OH>2.   In this  research, air was not used in mixing,
 and the mixing  time was  short, about 27 min.  Because of the hard
 alkaline nature  of  the tap water  and the conditions of mixing,  one
would  expect  the precipitate  to be mainly FeCOs; however,  the exact
nature was not determined.

Trickling filter effluent.  Final effluent from the Ames Water  Pollu-
 tion Control Plant was filtered.  The Ames sewage treatment plant
consists of comminutors, pumping, aerated grit chambers, primary
                                320

-------
settling, and standard rate trickling filters followed by final
clarifiers.

Secondary effluent from the final clarifiers was pumped from the
final collection chamber to the filters.  Most of this filtration of
trickling filter effluent was done during the month of July 1973.
The raw wastewater during July averaged 130 mg/1 BOD5, 161 mg/1 sus-
pended solids, and 337 mg/1 COD.  The secondary effluent for July
averaged 16 mg/1 BOD5, 20 mg/1 suspended solids, and 68 mg/1 COD.

Aluminum sulfate coagulated trickling filter effluent.  A pilot plant
was in operation at the Ames Water Pollution Control Plant that was
using alum to coagulate and flocculate secondary effluent for phos-
phorous removal.  The secondary effluent was pumped to an erdlator
(upflow solids contact unit) where approximately 200 mg/1 of alum
were added for precipitation of the phosphorous.

Average performance for the effluent from the erdlator during the
summer of  1973 was as follows  [118]:

                                         Influent     Effluent
     Average  suspended solids  (mg/1)       25.4        10.32
     Average  turbidity (units)             10.9         2.35
     Average  BOD5  (mg/1)                   38.6        10.80
     Average  TOG  (mg/1)                    14.0         7.59
     Average  total P04 (mg/1)              19.3         3.91
     Average  ortho P04 (mg/1)              18.7         2.39
     Average  total Kjehldahl N (mg/1)       4.9         4.93

The settled effluent  described above was used as  the  filter influent
for this study.

Activated  sludge  effluent.  An activated sludge pilot plant was  oper-
ated so  as to produce an  effluent to  filter.  The pilot plant was  a
Smith  and  Loveless  "Oxigest"  (Smith and Loveless  Model "C" Oxigest,
Smith  and  Loveless, Division-Union Tank Car  Company,  Lenexa, Kansas)
unit located  at the Ames  Water Pollution Control  Plant.  The aeration
tank volume is  2000 gal.,  and  the settling tank volume is 672 gal.

The  "Oxigest11 unit was operated at a  constant rate of 5 gpm.  This
gave a detention  time of  6.67  hr. The influent  to the unit was  pri-
mary effluent from the Ames Water Pollution  Control Plant.  The  aver-
 age characteristics of influent to the "Oxigest"  for the month of
July  1973  were  82 mg/1 of 5 day BOD,  65 mg/1 suspended solids, and
 204 mg/1 COD.  Typical characteristics of  effluent from  the "Oxigest"
unit were  5 to  15 mg/1 BOD5 and 9 to  10 mg/1 suspended solids.   The
mixed  liquor  suspended solids  was maintained at  around 2400 mg/1.
 Effluent from this activated  sludge pilot  plant was pumped from the
 effluent trough t6 the filters.
                                  321

-------
 Lime-soda ash softening precipitate.   The water  for  the  City  of Ames
 comes from ground water.  This  water  is  softened in  a modified, split
 treatment, lime-soda ash system.   The water  from wells is  aerated,
 slaked lime is added and mixed, and without  intermediate settling,
 the split flow and soda ash  are added and mixed  to precipitate addi-
 tional hardness.   The water  is  then allowed  to settle, and some of
 the sludge formed from the settling precipitants is  returned  to aid
 in the precipitation reactions  in the mixing step.   After  settling,
 approximately 2 mg/1 of polyphosphate is  added to stabilize the water
 and stop the reaction of chemical precipitation  so that  further pre-
 cipitate will not form on the filter  media,  piping,  or in  the clear
 wells.  Chlorine  and fluoride are also added.

 The pilot filter  plant was placed in  the  pipe gallery at the Ames
 plant, and water  was taken out  of the influent pipe  leading to one
 of the city filters.   The average turbidity  of the influent during
 this study was 6.4  FTU,  with a  range  of 4.3  to 12 FTU.   The turbidity
 is due mostly to  presence of calcium  carbonate particles.

 Sampling and Measurement

 Filter performance  was monitored  by observing head loss  development
 and filter influent and  effluent  quality.  Head  loss was measured,
 as previously discussed,  by head  loss buildup using mercury manom-
 eters.   Quality was measured by periodically collecting  a  grab
 sample and analyzing  it.   The period  between samples varied with
 the rate of  head  loss  buildup.  If head loss increased rapidly, the
 interval between  samples was decreased.  Samples were collected
 every  20 min for  iron  filtration  and up to every 3 hr for  softening
 precipitate  filtration.

 The quality  of  the  influent and effluent was measured in various
ways.  For the  iron filtration  series, the total iron in the sample
was measured.  For  the softening  precipitate, trickling  filter efflu-
 ent, activated  sludge  effluent, and the alum flocculated secondary
 effluent,  turbidity was used as one measure of quality.  Since sus-
pended solids are frequently specified as an effluent quality param-
 eter for  sewage treatment  plants, grab samples were taken and com-
posited  over  several time  intervals for suspended solids analysis.
When an  adequate amount of sample had been collected, a suspended
solids analysis was done.  Researchers have found that, within
 limits,  suspended solids concentrations found in treated wastewater
can be roughly correlated  to turbidity measurements  [59,128].   Never-
theless,  suspended solids  analyses were also run in this study be-
cause it was felt that the turbidity measurement might not measure
the larger particles in the filter influent and effluent.  For each
of the treated wastewaters filtered,  after the suspended solids were
filtered  from the sample,  a sample of the filtrate was taken and the
turbidity determined.  This result was considered to be background
color or colloidal material.
                                 322

-------
Iron measurement.  Iron was measured using 1,10 phenanthroline.  This
method is a spectrophotometric method in which the complex formed
with the ferrous iron produces an orange-red color that obeys Beer's
law.  The method was developed so that the analysis of iron samples
could be automated by using the Technicon Auto Analyzer II (Technicon
Industrial Systems, A division of Technicon Instruments Corporation,
Tarrytown, New York).  Forty samples could be analyzed per hour.
After the sample was taken, a few drops of hydrochloric acid and some
hydroxylamine hydrochloride were added to reduce the ferric iron to
ferrous iron and to acidify the mixture to keep the iron in solution.
By doing this, the samples were preserved for later analysis on the
automatic analyzer.

Turbidity measurement.  Various turbidimeters were used to determine
turbidity for this study.  As the quality differences are based on
the relative turbidity in the filter influent and filtrate, the ef-
fect of using different machines was negligible.  The instruments
used were all manufactured by Hach  (Hach Chemical Company, Ames,
Iowa) and included the Model 2100 and Model 2100A laboratory turbidi-
meters, the "Surface Scatter 3," and the Low Range Turbidimeter,
model 7120.  The  latter two instruments are continuous flow and
reading turbidimeters.  All the turbidimeters were calibrated  against
prepared  standards of  fonnazin polymer, prepared by Hach.  The unit
of  turbidity measurement was the Formazin Turbidity Unit  (FTU).

The continuous  flow  turbidimeters were checked  against the  laboratory
 turbidimeters.  Very close  agreement was  found.  The  continuous flow
 turbidimeters were used  to measure  the  influent to  the filters when
 filtering trickling  filter  effluent,  the  alum floe, and  the  softening
 precipitate.  The effluent  from  the filters was monitored by grab
 samples  analyzed  on the  laboratory  turbidimeters.

 Suspended solids  measurement.  The  procedure adopted  for the deter-
mination of total suspended matter  was  a slight modification of the
 procedure given in Standard Methods [117].   Whatman GF/C glass fiber
 filter paper was  used.  The filter  disks  were not prewashed and
 dried,  as it was known from prior experience with this  type of paper
 that the effect of not washing would be negligible.

                                 Results

 Quality and Head Loss

 In this empirical study, filter performance was measured by head loss
 buildup and effluent quality.

 The head loss buildup was measured in inches of mercury.  Quality
 was measured by various parameters, as previously described.  The
 effluent concentration divided by the influent concentration, C/C ,
 was then calculated.  Both head loss buildup and C/CO were plotted
 against the total volume of filtrate.  Total volume of filtrate
                                  323

-------
 during a filter run was  determined  by taking  the  flow rate, multi-
 plying it by  the time, and  summing  to yield the total volume of fil-
 tered water.   Total volume  of  filtrate was used to attempt to elimi-
 nate the influence  of any small  difference in flow rate that existed
 between the filters during  a filter run.

 As  previously described, five  different suspensions were filtered by
 two different dual  media; thus there  were ten series of filter runs.
 The series are identified in Table  49.  Several filter runs were made
 in  each series so that a consistent trend in  the data was observed.
 Some conditioning runs were undertaken at the beginning of each
 series in which the suspension was  filtered without close control
 to  acclimate  the filter media  to the  suspension.

 Table 49.  Identification of experimental series.


 Series   Dual media used            Suspension filtered


  I S    Sharp interface    Iron  floe

  I M    Mixed interface    Iron  floe

  II S    Sharp interface    Activated  sludge effluent

  II M   Mixed interface    Activated  sludge effluent

 III S    Sharp interface    Alum coagulated trickling filter eff

 III M   Mixed interface    Alum coagulated trickling filter eff

  IV S    Sharp interface    Trickling  filter effluent

  IV M   Mixed interface    Trickling  filter effluent

  V S    Sharp interface    Lime-soda  ash softening ppt

  V M   Mixed  interface    Line-soda  ash softening ppt
A typical plot of head loss buildup versus volume of filtrate and
C/C0 versus volume of filtrate for each series is presented in Figs.
100 through 109.

There are three different parameters presented in the graphs of head
loss versus volume of filtrate.  They are (1) the head loss buildup
in the combined media (that media which was in column 1, Fig. 99)
labeled "COMBINED TOTAL HL," (2) the head loss buildup in the sepa-
rate  coal (that media which was in column 2, Fig. 99) labeled
"SEPARATE COAL HL," and (3) the sum of the head loss buildup in the
separate coal and separate sand (that media which was in column 2
                                 324

-------
O)
Q
    0.80-
    0.60
    0.40
    0.20
    0.00
    4.00
    3.00
2.00
    1.00
    0.00
          QUALITY
       O  COMBINED C0
       A  SEPARATE C/C0
          HEAD LOSS (HL)
        O COMBINED TOTAL HL
        A SEPARATE COAL HL
        Q SEPARATE TOTAL HL
        0.00   8.00   16.00   24.00   32.00   40.00   48.00 56.00

                       VOLUME FILTRATE, gal. (x 101)
  'Fig. 100.  Head loss and filtrate quality vs volume of filtrate,
             Series I S, run 2,  sharp Interface,  filtration of
             iron with C  - 8.68  to 9.64 mg/1 Fe, Avg 9.07 mg/1.
                       o
                               325

-------
  0.16
  0.12
 o
u
  0.08
  0.04
  0.00
  8.00
o>
I
.c'6.00
 H

O
o4.00

x

  2.00



  0.00
             QUALITY
          o  COMBINED C0
          A  SEPARATE C/C0
              HEAD LOSS (HL)
           O  COMBINED TOTAL HL
           a  SEPARATE COAL HL
           D  SEPARATE TOTAL HL
                                                              I
      0.00   4.00     8.00    12.00    16.00    20.00   24.00    28.00
                     VOLUME FILTRATE, gal. (x 101)
   Fig. 101.  Hetd  lomm and filtrate quality v« volua* of filtrate,
             S*ri«« I M, run 11,  mixed interface,  filtration of
             iron  with C0 - 9.2 to 9.7 mg/1 Fe, Avg  9.42 mg/1.
                               326

-------
0.80
0.60
0.40
0.20
0.00
  8.00

o>
x
.c'6.00
 s

9
Q 4.00

x

  2.00
 0.00
            QUALITY
          o COMBINEDC/CO
          A SEPARATE C/C0
           HEAD LOSS (HL)
        O  COMBINED TOTAL HL
        a  SEPARATE COAL HL
        D  SEPARATE TOTAL HL
     0.00   8.00    16.00   24.00   32.00   40.00   48.00   56.00

                    VOLUME FILTRATE, gal. (x 101)
 Fig.  102.  Head Loss  and filtrate quality vs volume of  filtrate,
           Series II  S, run 1, sharp  interface, filtration of
           activated  sludge effluent  with Co = 4.5 to 8.5 FTU,
           Avg 6.16 FTU.
                              327

-------
          QUALITY
        O COMBINED C/Co
        A SEPARATE C/Co
             HEAD LOSS (HL)
          o  COMBINED TOTAL HL
          A  SEPARATE COAL HL
          D  SEPARATE TOTAL HL
   0.80
   0.60
  0.40
   0.20
  0.00
8.00
£6.00-
2
Q 4.00
  2.00
  O.OOLjQ,
      0.00   8.00   16.00   24.00   32.00   40.00   48.00   56.00

                      VOLUME FILTRATE, gal. (x 101)


   Fig. 103.  Head loss and filtrate quality vs volume of filtrate,
             Series  II M, run 1, mixed  interface,  filtration of
             activated sludge effluent  with Co - 2.6  to  8.6 FTU,
             Avg 4.15 FTU.
                               328

-------
           QUALITY
        o  COMBINED C/Co
           SEPARATE C/C0
           HEAD LOSS (HL)
        O COMBINED TOTAL HL
           SEPARATE COAL HL
        D SEPARATE TOTAL HL
0.00
    0.00    8.00    16.00    24.00    32.00    40.00   48.00   56.00

                    VOLUME FILTRATE, gal. (x 101)


  Fig. 104.  Head loss and filtrate quality vs volume of filtrate,
            Series III S, run 3,  sharp interface,  filtration of
            alum coagulated trickling filter effluent with C0 -
            3.8 to 10 FTU, Avg 5.21 FTU.
                              329

-------
  QUALITY
o COMBINEDC/CO
A SEPARATE C/C0
   HEAD LOSS (HL)
o  COMBINED TOTAL HL
A  SEPARATE COAL HL
D  SEPARATE TOTAL HL
    0.80



    0.60



    0.40



    0.20



    0.00



    8.00



.E   6-°°
 s

2
Q   4.00

i

    2.00



    0.00
       0.00   5.00   10.00   15.00   20.00   25.00   30.00 35.00

                      VOLUME FILTRATE, gal. (x 101)

  Fig. 105.  Head loss and  filtrate quality vs volume of  filtrate,
            Series III M,  run 5, mixed interface, filtration of
            alum coagulated trickling filter effluent with Co -
            1.7 to 6.2 FTU, Avg 2.95 FTU.
                   330

-------
    o
    00
    o
    •o
    QUALITY
O   COMBINED C/C
A   SEPARATE C/C  °
    O !
      I
      ]
    si
     • i
    o
   8
     •
   00
 oo
 q  s
    8
    CM*

    8
     •
    o
    HEAD LOSS (HL)
O   COMBINED TOTAL HL
A   SEPARATE COAL HL
D   SEPARATE TOTAL HL
      0.00    5.00    10.00   15.00   20.00   25.00   30.00   35.00
                       VOLUME FILTRATE, gal. (xlO1)
Fig. 106.   Head  loss and filtrate quality vs volume of  filtrate,
           Series  IV S, run 3, sharp interface,  filtration of
           trickling filter effluent with Co = 4.7 to 8.5 FTU,
           Avg 6.29 FTU.
                               331

-------
 0.80
U
 0.60
 o
 0.40
 0.20
   0.00



   8.00

 o>
X
 c 6.00



|,«,

i

   2.00
 0.00
                                           QUALITY
                                        O COMBINED C/C
                                        A SEPARATE C/C  °
                          HEAD LOSS (HL)
                       O COMBINED TOTAL HL
                       A SEPARATE COAL HL
                       D SEPARATE TOTAL HL
    0.00    4.00    8.00   12.00   16.00   20.00    24.00   28.00
                   VOLUME FILTRATE, gal.  (xlO1)
 Fig.  107.  Head loss and filtrate quality vs volume of filtrate,
           Series IV M, run 3, mixed interface, filtration  of
           trickling filter effluent with CQ - 12 to 15 FTU,
           Avg 6.14 FTU.
                             332

-------
                                            QUALITY
                                          o COMBINED C/C0
                                          A SEPARATE C/C0
                                           HEAD LOSS (HL)
                                        O  COMBINED TOTAL HL
                                        £  SEPARATE COAL HL
                                        o  SEPARATE TOTAL HL
  0.00
                                      I
I
                     8.00    12.00    16.00   20.00    24.00  28.00

                     VOLUME FILTRATE, gal. (x 10*)
Fig.  108.  Head !••• and filtrate quality YB veluae of filtrate,
          Sorie* Y S, run 4, •harp laterfact, filtration of lina-
          •«4a aah aoftaming pvaclpitata with C0 • 5.6 to 6.5
          RU, *rg 6.14 FTU.
                             333

-------
   0.00
                                               QUALITY
                                            o  COMBINED C/Co
                                            A  SEPARATE C/Co
                                            HEAD LOSS (HL)
                                          O COMBINED TOTAL HL
                                          A SEPARATE COAL HL
                                          a SEPARATE TOTAL HL
                                    A	A
1
       0.00   2.00    4.00     6.00     8.00    10.00    12.00  14.00

                      VOLUME FfLTRATE, gal. (x 102)
Fig.  109.  Head loss end  filtrate quality vs volume of filtrate,
          series V M,  run 3, mixed Interface, filtration of lime-
          soda ash softening precipitate vlth Co - 6.8 to 8.3 FTU,
          Avg 7.5 FTU.
                              334

-------
and column 3, Fig. 99, respectively) labeled "SEPARATE TOTAL HL."

The graphs of C/C0 versus volume of filtrate compare the effluent
from the combined media (column 1, Fig. 99) labeled "COMBINED C/CO"
to the effluent from the separate media (the effluent from column 3,
Fig. 99) labeled "SEPARATE C/CO."

Suspended Solids

Suspended solids tests were run on the three treated wastewater
streams, both influent to and effluent from the filters.  The range
of values and the average value for the influent, combined media
effluent, and the separate media effluent are given in Table 50 for
all suspended solids data from the indicated series.

Table 50.  Influent and effluent suspended solids data, average and
           range, for wastewater series.
Influent Series
suspension no.
II S
Activated
sludge
effluent II M

Alum m S

coag
TF
effluent IIZ M
IV S
Trickling
filter
effluent IV M



Range
Avg

Range
Avg
Range
Avg

Range
Avg
Range
Avg

Range
Avg
Influent,
mg/1
4.8-8.0
6.0

3.8-14.6
9.53
5.0-10.2
8.05

5.6-11.2
7.61
10.0-30.8
19.37

18.0-34.8
23.30
Combined eff,
mg/1
0.9-2.6
1.38

0.3-1.2
0.60
1.0-3.9
2.05

0.5-2.9
1.49
3.0-9.6
5.50

1.9-6.2
4.48
Separate eff,
mg/1
0.9-2.3
1.26

0.2-1.2
0.61
1.0-3.2
1.91

0.4-2.8
1.36
3.0-9.0
5.57

2.0-6.2
4.13
Background Turbidity

The background turbidity  (turbidity of filtrate remaining from the
suspended solids tests) of the three treated wastewater streams was
determined.  The average  values of the background turbidity are
given  in Table 51.  From  the data presented, it can be seen that the
background turbidity was  affected very little by filtration (the
background turbidity of the influents was approximately equal to the
                                  335

-------
Table 51.  Average values of background (bg) turbidity for the
           treated wastewaters.
    Influent
   suspension
Influent bg
turbidity,
    FTU
             Combined eff
            bg turbidity,
                 FTU
 Separate eff
bg turbidity,
     FTU
Activated sludge

Alum coagulated
TF effluent

Trickling filter
effluent
0.56


0.53


1.60
                    0.59
                    0.59
                     1.70
     0.59


     0.58


     1.80
background of the effluents).  It may also be noted that the back-
ground turbidity was a significant percentage of each total effluent
turbidity for these treated wastewater streams.  Furthermore, this
colloidal background matter is not measured by the suspended solids
analysis.

                              Discussion

The results presented in Figs. 100 through 109 are summarized in
Table 52.  Where there was no difference or only a slight difference,
"no difference" is indicated.  Where the separate media had a better
performance (i.e., lower head loss buildup or better filtered water
quality), "separate better" is indicated.  In the series where the
combined media had a better performance, "combined better" is indi-
cated.  In Series II S and Series IV S, no comparison could be made,
as will be explained later; therefore, "no comparison" is indicated.

Table 52.  Summary of results comparing filter performance for sharp
           and mixed interface media.
Series   Filtering   Interface
                Head loss
                                Quality
I S
I M
II S
II M
III S
III M
IV S
IV M
V S
V M
Fe
Fe
AS
AS
Alum
Alum
TF
TF
Soft
Soft
Sharp
Mixed
Sharp
Mixed
Sharp
Mixed
Sharp
Mixed
Sharp
Mixed
Separate better
Separate better
No comparison
No difference
Combined better
Separate better
No comparison
Separate better
No difference
No difference
No difference
No difference
No comparison
No difference
No difference
No difference
No comparison
No difference
Separate better
Separate better
                                 336

-------
The sharp interface media was used as a control to establish what
differences in performance could be attributed to the equipment and
operating procedure.  Thus, if the sharp interface media presented a
significant trend in one direction and the mixed interface media
demonstrated the same trend in greater magnitude or the opposite
trend this difference could be attributed to the mixing of the media
at the interface.

Upon examination of the data and plots for the filtration of the iron
floe, it can be seen that there is no difference in the quality that
can be attributed to media mixing at the interface.  The head loss
data and plots reveal that the mixing at the interface resulted in
slightly larger head loss.

No valid conclusions can be drawn from the filtration of the acti-
vated sludge effluent or the trickling filter effluent.  This is be-
cause all the removal was in the coal during sharp interface filtra-
tion.  Thus, the control offered by the sharp interface data was
lost, and the effect of the equipment and operating procedure was
not known.

The data and plots for the filtration of the alum coagulated trick-
ling filter effluent by the sharp interface media showed the equip-
ment and operating differences caused the combined media to have
less head loss buildup than the separate media.  However, the data
and plots of the filtration of this suspension by the mixed interface
dual media show that the separate media have less head loss buildup.
Thus, the effect of the intermixing at the interface was a detrimen-
tal one because there was a greater head loss in the combined mixed
interface media.  This greater head loss is attributable to the
intermixing.  The quality data and plots could show no difference in
quality attributed to intermixing.

The filtration data for the lime-soda ash softening precipitate re-
vealed no difference in head loss buildup between the combined and
separate media for both the mixed and sharp interface dual media.
The quality data and plots for the sharp interface media showed that
the separate media had a better effluent.  This observed difference
can be attributed to the equipment and operating procedure.  The
quality data and plots for the mixed interface media showed that the
separate media also produced a better quality effluent.  Since both
the sharp and mixed interface media had the same trend, no difference
in quality can be attributed to the mixing of the dual media at the
interface.  A visual observation of the upper surface of the separate
sand revealed a  layer of white precipitate.  This blanket of precipi-
tate on the upper surface of the separate sand in both the sharp and
mixed interface media may have acted like a filter itself, thereby
improving the effluent quality from the separate media.
                                 337

-------
 From the  results,  it can be  seen  that  the mixing of the dual media
 at the  interface causes little  or no effect on either head loss
 buildup or  effluent quality  of  a  dual-media filter.  Some slight in-
 crease  in head  loss buildup  may be attributed to media mixing at the
 interface for some suspended solids, but this is not true of all
 solids, and the magnitude  is slight.

 Filtrate  quality as measured by turbidity or iron concentration
 showed  no benefit  or detriment  due to  intermixing.  The suspended
 solids  data in  Table 50 also show no apparent effect of the media
 intermixing at  the interface.

 A point must be made concerning the effluent quality.  As can be
 seen from the data, all of the  suspended material was not completely
 removed (C/CO greater  than zero).  Thus, some of the suspended matter
 must have reached  the  interface if some passed through the entire
 dual-media  filter.  In the case of Series II S and IV S, little of
 the  suspended matter that  reached the  interface or the sand was re-
 tained  due  to the  size of  the sand.

 It would  be of  interest to compare these observations with theoreti-
 cal  models  of filter performance.  However, it is difficult to draw
 firm conclusions because the models have been derived and verified only
 for  single-media filters.  Nevertheless, an attempt will be made to
 apply filtration theory to support the observations.  Boyd and Ghosh
 [18]  recently summarized the various models for filtrate quality.
 They restated the now  commonly  accepted idea that removal of sus-
 pended  solids by a granular  filter involves a transport step and an
 attachment  step.

 All  of  the  parameters  in the various models for both transport and
 attachment  are  properties  of the particles in suspension or the
 carrying  fluid  except  for  the diameter of the filter grains.   Two of
 the  transport models,  namely, interception and Brownian diffusion,
 indicate  that removal  effectiveness is an inverse function of grain
 diameter.    Therefore,  one would expect better filter efficiency with
 finer grain size,  if all other  parameters were unchanged.

Most models  for prediction of head loss in granular media filtration
 formulate head  loss to be  a  function of solids capture [64,84],  Thus
 if the  solids capture  is the same for  the combined or separate media,
 the  total head  loss would be expected  to be about the same.

The models  do not deal with  the effect of grain diameter in the in-
termixed zone on the efficiency of a dual-media filter.   If the same
grains are present they should yield the same removal efficiency
whether they are intermixed  in  a dual-media bed,  or operated sepa-
rately  in series.  This idea would support the experimental results
reported herein which  showed no difference between the filtrate for
the combined and separate media.  However,  it is  probable that the
grain diameter in the models is really important  as an indirect
                                  338

-------
measure of the size and shape of the void spaces in the filter bed.
When the small sand grains move up into the coarser coal layers,
they tend to fill the larger voids of the coal, causing a pore size
and permeability between that associated with the sand and the coal
as shown previously in Chapter X.  Similarly, the coal grains that
subside into the upper sand layer create a pore size and permeability
between the sand and coal.  Intermixing resulted in a total bed depth
slightly less than the sum of the depths of the two media, measured
individually in this work.  A similar observation had also been re-
ported previously in Chapter X.  Since the overall depth of the dual-
media bed is less, it would have a lower clean bed porosity and per-
meability in the intermixed zone.  Because of this, one would expect
better filtrate quality but higher head loss from the combined media
than the separate media.  The higher head loss for the combined media
was observed for three of the solids reported herein, but the better
quality was not observed.

Substantial differences in performance were observed between the
sharp and mixed interface media.  These differences would be expected
from the foregoing theories.  The mixed interface media produced a
better filtrate quality than the sharp interface media, but also
generated higher head loss.  Substantially the same size coal was
used in the two media, but a much finer sand was used in the mixed
interface media than in the sharp interface media.  From the theory
presented, the finer sand should result in more efficient solids re-
moval, which in turn would generate more head loss.  These media
sizes were purposely selected to achieve the desired degree of inter-
facial intermixing or separation.  The coarse sand in the sharp in-
terface media was required to produce a sharp interface with the ex-
pectation that it would cause poorer solids capture and lower head
loss.

The degree of interfacial intermixing is dependent on the size, den-
sity, and shape of adjacent media, as well as the backwash flow ve-
locity and valve shut-off procedure as shown previously in this re-
port in Chapter X.  If a normal coal size and gradation are selected,
one cannot achieve a sharp interface except by sacrificing expected
filtrate quality.

The controversy as to the desirability or detriment of interfacial
intermixing was started by Conley  [32] and Camp  [23],  Camp  [23,24]
stated that he felt a sharp interface acts as a safeguard against
breakthrough.  He discouraged  intermixing.   Ives  [63] and Gregory
 [54] have also implied that one  should discourage  interfacial  inter-
mixing.  However, other researchers and actual case studies have
shown that filters perform quite adequately when intermixing occurs
at the interface either by accident or by plan.

The results of this study showed substantially better filtrate qual-
ity for  the mixed interface media  than for the  sharp  interface media.
                                  339

-------
Thus the results are in opposition to the design recommendation of
Camp.  The better filtrate observed for the mixed interface media is
due to the use of finer sand and not to the intermixing itself.
However, intermixing is the inevitable result of selecting the dual
media to provide substantially finer top sand size than top coal
size.

None of the researchers have presented proof to back up claims that
the intermixing itself is good or bad.  Brosman and Malina  [19] are
the only ones who have attempted to measure the effect of inter-
mixing on both head loss and effluent quality.  The writers feel that
their study had some experimental weaknesses  [109] which led them
to the false conclusions that interfacial intermixing, in and of it-
self, provided benefits to both filtrate quality and head loss.  That
conclusion is not supported by the results reported herein.

                              Conclusions

The following points are concluded from this study:

1.   Interfacial intermixing does not in itself affect filter per-
     formance as measured by both head loss development and effluent
     quality of dual-media filters.

2.   The mixed interface filter media produced better filtrate qual-
     ity than did the sharp interface filter media for all suspen-
     sions filtered in this study; this was due to substantially
     finer sand in the mixed interface media.

3.   The mixed interface filter media produced higher head losses
     than the sharp interface filter media for all suspensions due
     to greater suspended solids removal.

4.   Intermixing at the interface of dual-media filters is an un-
     avoidable phenomenon which results when United  States anthracite
     (sp.  gr.  about 1.7) and sand are used and the sizes are selected
     to achieve coarse to fine filter media in the direction of flow.
                                 340

-------
             XII.  ABRASIVE LOSS OF COAL DURING AIR SCOUR
The effect of air scour on anthracite coal filter media became of
concern because of the possible widespread use of air scoured fil-
ters using anthracite coal as one of the filter media.  Widespread
use seems probable for the following reasons:   (1) the apparent bene-
fits of air scour as an adjunct to water backwash, (2) the apparent
necessity of auxiliary agitation in wastewater filtration, (3) the
necessity of using dual media in wastewater filtration to achieve
adequate filter run length.

The visual appearance of coal filter media when subjected to air
scour was observed in plastic filter housings.  The coal appears to
be in rather violent motion, especially the coal near the top surface
of the bed.  The granules of the coal are in contact with each other
and rubbing against each other, unlike those in a fluidized bed where
collisions and rubbing are absent.  Thus the abrasive loss of coal
was of concern due to the softer nature of coal as compared to other
common filter media such as silica sand and garnet sand.

An experiment was conducted to gain some view of the potential seri-
ousness of the abrasive loss problem.  The experiment consisted of
subjecting a filter bed of anthracite coal to essentially continuous
air scour for two weeks.  Two weeks of such exposure would be com-
parable to about 20 years of normal filter operation assuming 24-hr
filter cycles and 3 min of air scour per cycle.  The filter was
water backwashed twice a day to observe the visual loss of coal dust.
Furthermore, changes in the depth of bed and pressure loss profile
through the bed in both upflow and downflow were observed to obtain
evidence of changes in the filter media.

                        Experimental Procedure

The details of the experiment were as follows:

The starting coal was a standard crushed anthracite coal "Philterkol"
obtained from the Reading Anthracite Coal Company, Pottsville, Penn-
sylvania.  The suppliers specified effective size was 0.6 to 0.79 mm,
the uniformity coefficient was less than 1.8,  and the reported hard-
ness was 3 on the Moh Scratch Test comparison scale.

The coal was split to obtain a representative sample, and 16 in. of
coal was placed in a 6-in. ID pilot filter.  The coal was backwashed
with water and allowed to stratify.  A total of about 2 in. of the
fine coal was skimmed from the surface of the stratified bed in two
increments during backwashing.  The coal was then removed from the
bed, air dried and oven dried, and two separate representative sam-
ples were collected for sieve analysis.  A riffle type splitter was
used to collect the representative sample.
                                 341

-------
 The  sieve analyses were performed on a nest of United States Standard
 sieves  (12,  14,  16,  18, 20, 25, and 30 mesh) using a Tyler portable
 shaker  for 5 min.  The sample size was 356.6 and 407.9 g for the two
 samples.

 The  remainder of the coal was weighed and was placed back in the fil-
 ter  column,  fluidized to permit stratification and the pressure loss
 profile was  observed from the piezometer tubes located at 3-in. depth
 increments as shown  in Fig. 4.  A water temperature of 22 °C  was used.

 After the upflow and downflow pressure profile observations, the
 water level was  lowered to about 6 in. above the coal surface, and
 the  air scour was applied at a rate of 4 cfm/sq ft.  The air was kept
 on continuously except for two short periods each day when the air
 was  shut off and the filter was subjected to backwash with 50% bed
 expansion, and to upflow and downflow pressure profile observations.
 The  amount of coal dust in the initial backwash water was observed
 qualitatively as one measure of abrasive loss.  The downflow pressure
 drop between piezometer taps at 7 gpm/sq ft and 22 °C was  observed  to
 see  if  there was any progressive increase in pressure drop due to the
 formation of finer media size as a result of abrasion.   The pressure
 drop during upflow at a rate which would provide 50% bed expansion at
 22 °C was measured  to see  if  there were any  progressive changes due
 to abrasive  loss.  The upflow rate was selected to provide an ex-
 panded  bed depth of 21 in.  The development of a finer media would
 result  in greater expansion at a given upflow rate and less pressure
 drop between the piezometer taps.  The depth of the bed after each
 backwash was observed after carefully and slowly closing the backwash
 valve to ensure a consistent degree of packing of the bed.  Any major
 loss of media or change in sphericity due to abrasion would manifest
 itself  by a decrease in bed depth.

 At the  end of two weeks of the above procedure, the coal was removed
 from the filter, dried, weighed, and again two representative samples
 were taken and subjected to sieve analysis.

                                Results

 Some abrasive loss was evident from the color of the coal dust in the
water at the end of each period of air scour.  The initial backwash
 water was black with coal dust but cleared up in the first 2 min of
 backwash.

 There was no substantial loss in bed depth or bed weight as shown in
Table 53.   Furthermore, there was no substantial change in head loss
 either upflow or downflow, indicating no  substantial change in size
 of the media.

The sieve analysis  before and after the two-week period of air scour
 is presented in Fig.  110.   Each curve represents the average of two
                                 342

-------
Table 53.  Changes in coal bed over a two-week period of air-scour
           exposure equivalent to about 20 years of normal service.

Bed depth,
Bed weight

in.
dry, g
Initial
value
14-1/8
5019
After
two weeks
13-7/8
4775
Total downflow head loss, ft

Upflow rate to achieve 21 in.
0.30
0.29
bed depth, gpm
Total upflow head loss, ft H20
Effective size of media, mm
Uniformity coefficient
5.1
0.35
0.80
1.59
5.3
0.34
0.78
1.60
sieve analyses.  It is evident that the coal decreased in size
slightly in this air scour exposure, about 0.02 mm in effective size,
which is a negligible change.

                              Conclusions

A typical crushed anthracite coal filter media was subjected to air
scour for two weeks to simulate the abrasive effects of 20 years of
typical filter service.  From this study, it is concluded that some
abrasive loss of coal does occur, but it is a negligible amount.  The
total loss of media and changes in media head loss and size were less
than 3%.  This is considered negligible because coal filter media
suppliers usually request and are granted about a 107o tolerance on
the effective size of the media which they supply.
                                  343

-------
N
CO
    1.8
    1.6
    1.4
    1.2
    1.0
    0.8
    0.6
    0.4
             7-12-72
                BEFORE
                AIR SCOUR
             CD
AFTER
AIR SCOUR
       12     5    10    20  30  40  50 60  70   80     90    95
                  PERCENT PASSING BY WEIGHT, % finer
   Fig. 110.  Sieve analysis of coal before and  after  14 days of
              continuous air scour.
                                344

-------
                      XIV.  REFERENCES
1.  Adler, I. L., and Happel, J.  The fluidization of uniform smooth
    spheres in liquid media.  Chem. Eng. Prog., Symp. Ser., 58(38):
    98-105, 1962.

2.  American Society for Testing Materials.  1968 Book of ASTM
    Standards.   Part 10.  Concrete and mineral aggregates.  ASTM,
    Philadelphia, 1972.

3.  American Water Works Association, Inc.  Water Quality and Treatment.
    New York, McGraw-Hill Book Co., Inc., 1971.

4.  Amirtharajah, A.  Expansion of graded sand filters during backwash-
    ing.  Unpublished M.S. thesis.  Ames, Iowa, Library, Iowa State
    University of Science and Technology, 1970.

5.  Amirtharajah, A.  Optimum expansion of sand filters during back-
    wash.  Unpublished Ph.D. thesis.  Ames, Iowa, Library, Iowa
    State University of Science and Technology, 1971.

6.  Amirtharajah, A. and Cleasby, J. L.  Predicting expansion of
    filters during backwashing.  J. Am. Water Works Assoc., 64:52-59,
    1972.

7.  Anderson, T. B., and Jackson, R.  The nature of aggregative and
    particulate  fluidization.  Chem. Eng. Sci., 19:509-511, 1964.

8.  Babbitt, H.  E., Doland,  J. J., and Cleasby, J. L.  Water Supply
    Engineering. 6th ed.  New York, McGraw-Hill Book Co., Inc., J.yb2.

 9.  Barret, A. D.   Immedium  filter in tertiary treatment.  Process
    Biochem., 6_(1): 33-35,  1971.

10.  Baylis, J. R.   Nature  and effects of  filter backwashing.  J. Am.
    Waterworks  Assoc.,  51:126-156, 1959.

11.  BayUs, J. R.   Review  of filter design and methods of washing.  J. Am.
    Water Works  Assoc.,  51:1433-1454, 1959.

12.  Becker, H. A.   An investigation of  the  laws governing  the spouting
    of coarse  particles.   Chem. Eng. Sci., 13:245-262, 1961.

13.  Boby, W. M.  T.,  and Alpe, G.   Practical  experiences using upward
     flow filtration.   Proc.  Soc.  Water  Treat.  Exam.,  16:215-229,  1967.

14.  Boss, R.  R.  Hydraulics and  intermixing of  sand and  coal  filters.
     Unpublished M.S.  thesis. Ames,  Iowa, Library, Iowa  State University
     of Science and  Technology,  1973.

15.   Botterill,  J. S. M.  Progress in fluidization.  Br. Chem.  Eng.,
     10:26-30, 1965.

                                   345

-------
 16.  Botterill, J. S. M.  Progress in fluidization.  Br. Chem. Eng.,
     13:1121-1126, 1968.

 17.  Bourgeois, P., and Grenier, P.  The ratio of terminal velocity to
     minimum  fluidizing velocity for spherical particles.  Can. J. Chem.
     Eng., 46:325-328, 1968.

 18.  Boyd, R. H., and Ghosh, M. M.  An investigation of the influences
     of  some  physiochemical variables on porous-media filtration. J.
     Am. Water Works Assoc., .66:94-98, 1974.

 19.  Brosman, D. R., and Malina, J. E.  Intermixing of dual media  filters
     and effects on performance.  Technical Report EHE-72-4 CBWR-86.
     Center for Research in Water Resources.  The University of Texas
     at Austin, 1972.

 20.  Buevich, Yu. A., and Markov, V. G. Pseudo-turbulent diffusion of
     particles in homogeneous suspensions.  Int. Chem. Eng., 10:570-
     575, 1970.

 21.  Cairns, E. J., and Prausnitz, J. M.  Longitudinal mixing in
     f luidization.  AIChE J., ji:400-405, I960;

 22.  Cairns, E. J., and Prausnitz, J. M.  Macroscopic mixing in fluidi-
     zation.  AIChE J., i:554-560,1960.

 23.  Camp, T. R.  (Discussion of) Experience with anthracite-sand
     filters.  J. Am. Water Works Assoc., 5.3:1478-1483, 1961.

 24.  Camp, T. R.  Theory of water filtration.  ASCE J. Sanitary Eng. Div.,
     90:1-30, 1964.

 25.  Camp, T. R.  (Discussion of) Theory of water filtration.  ASCE J.
     Sanitary Eng. Div., £1:55-69, October 1965.

 26.  Camp, T. R., Graber, S. D., and Conklin, G. F.  Backwashing of
     granular water filters.  ASCE J. Sanitary Eng. Div., No. SA6,
     pp. 903-925, December 1971.

27.  Camp, T. R., and Stein, P. C.  Velocity gradients and internal
     work in fluid motion.  J. Boston Soc. Civ. Eng., 30:219-237, 1943.

28.  Carvalho, W. J.  Fluidization of crushed anthracite coal.  Unpub-
     lished M.S. thesis.  Ames, Iowa, Library, Iowa State University of
     Science and Technology, 1970.

29.  Clark, J. W., Viessman, W., and Hammer, M. J.  Water Supply and
     Pollution Control.  Scranton, Pennsylvania, International Textbook
     Company,1971.
                                    346

-------
30.  Cleasby, J. L.  Iron removal,  a case study.   J. Am. Water Works
     Assoc., £7(3):147-149.   March  1975.

31.  Cleasby, J. L.,  and Baumann, E. R.   Selection of  sand  filtration
     rates.  J. Am. Waterworks Assoc.,  54(5):579-602,  1962.

32.  Conley, W. R.  Experience with anthracite-sand filters.   J.  Am.
     Waterworks Assoc., 53:1473-1478, 1961.

33*  Coulson, J. M.,  and Richardson, J. F.  Chemical Engineering. Vol.
     2, 2nd ed. New York, Pergamon Press Inc., 1968.

34.  Gulp, R. L., and Gulp, G. L.  Advanced Wastewater Treatment. New
     York, Van Nostrand Reinhold Company, 1971.

35.  Davidson, J. F., and Harrison, D.  Fluidized Particles.   New York,
     The Syndics of the Cambridge University Press, 1963.

36.  Davies, L., and Richardson, J. F.  Gas interchange between bubbles
     and the continuous phase  in a  fluidized bed.  Trans. Inst. Chem.
     Eng., 44:T293-T305, 1966.

37.  Degremont, S. A.  Water Treatment Handbook.  3rd English ed.
     Kensington, London W14, England, Hugh K. Elliot, Ltd., 1965.

38.  Dvorin, R.  Advances make wastewater filtration economical and
     effective  solution.  Water Sewage Works  116.(9) -.1W/12-1W/16, 1969.

39.  Efremor, G. I.  and Vakhrusher,  I.  A.  A study of the hydrodynamics
     of  three-phase fluidized  beds.   Int. Chem. Eng., 1():37-41, 1970.

40.  Ergun,  S.  Fluid  flow  through  packed columns.  Chem. Eng. Prog.,
     48:89-94,  1952.

41.  Evans,  S.  C., and Roberts, F.  W.  Twelve months' operation of
     sand  filtration and microstraining  at  Luton.   Inst. Sewage
     Purif., J. Proc., 4:333-341, 1952.

42.  Evans,  S.  C., and Roberts, F.  W.  Recent developments in sewage
     treatment at  Luton.  Inst. Sewage Purif., J.  Proc., 3_:225-236, 1955.

43.  Evers,  R.  H.  Tool up  with mixed-media filters.  Water Wastes Eng.,
     8(5):C14-C16, 1971.

44.  Ewing,  G.  W.  Instrumental Methods  of  Chemical Analysis.  3rd ed.
     New York,  McGraw-Hill  Book Co.,  Inc.,  1969.

45.  Fair, G.  M.,  and  Geyer,  J. C.  Water Supply an«j W,as,t;q Water Dis-
      posal.  8th  ed. New York, John Wiley and Sons,  Inc.,  1958.

46.   Fair, G.  M.,  Geyer, J. C.,  and Okun, D.  A.   Water and Wastewater
     Engineering.  Vol.  2.   New York, John  Wiley and Sons, Inc., 1968.
                                    347

-------
47.  Fair, G. M., and Hatch, L. P.  Fundamental factors governing the
     streamlike  flow of water through sand.  J. Am. Water Works Assoc.,
     25:1551-1565,  1933.

48.  Frantz, J.  F.  Minimum fluidization velocities and pressure drop
     in fluidized beds. 1 Chem. Eng., Symp. Series, 62(62):2l-31, 1966.

49.  Friedlander, S. K., and Topper, L.  Turbulence. Classic Papers
     on Statistical Theory.  New York, Interscience Publishers Inc.,
     1961.

50.  Furukawa, J., and Ohmae, T.  Liquidlike properties of fluidized
     systems.  Ind. Eng. Chem., 50:821-828, 1958.

51.  Galloway, T. R., and Sage, B. H.  A model of the mechanism of
     transport in packed, distended and fluidized beds, Chem. Eng. Sci.,
     25:495-516, 1970.

52.  Ghosh, M. M., O'Connor, J. T., and Engelbrecht, R. S.   Precipitation
     of iron in aerated ground waters.  ASCE J. Sanitary Eng. Div.,
     90(SA1):199-213, 1966.

53.  Godard, K., and Richardson, J. F.  Correlation of data for minimum
     fluidizing velocity and bed expansion in particulately fluidized
     systems.  Chem. Eng. Sci., 24; 363-367, 1969.

54.  Gregory, R.  Media specification and backwashing - W.R.A. Paper
     presented at Anthracite-Sand Filtration, A Working Conference,
     University of Reading, England, January 3-5, 1972.

55.  Handley, D., Doraiswamy,  A.,  Butcher,  K. L., and Franklin, N. L.
     A study of the fluid and particle mechanics in liquid-fluidized
     beds.  Trans. Inst. Chem.  Eng., 44:T260-T273, 1966.

56.  Hanratty,  T. J.,  Latinen,  G.  and Wilhelm,  R. H.  Turbulent
     diffusion in particulately fluidized beds  of particles.   AIChE J
     1:372-380,  1956.

57.  Harrison, D.,  Davidson, J.  F., and  Kock, J.  W.   On the nature of
     aggregative and particulate fluidization.   Trans.  Inst.  Chem.
     Eng., .39:202-211,  1961.

58.  Holding, J.  C.  Polishing  sewage effluents,  Effluent Water Treat.
     J., .12:665-667, 669-672, 1972.

59.  Huang, J. Y. C.  Granular  filters for  tertiary  wastewater treat-
     ment. Unpublished  Ph.D. dissertation.   Ames, Iowa, Library,  Iowa
     State University of Science and Technology,  1972.

60.  Hudson,  H. E.   Filter  backwashing experiments at the Chicago
     experimental plant.  J. Am. Waterworks  Assoc.,  27_:l5b7-l56b,  1935.
                                   348

-------
61.  Huisman, L.  Rapid filtration,  Part 1.  Mimeographed  book of
     lectures in English,  Delft University of  Technology,  Depart-
     ment of Civil Engineering, 1974.

62.  Hulbert, R.,  and  Herring,  F. W.  Studies  on the washing of rapid
     filters.  J.  Am.  Water Works Assoc.,  21:1445-1513, 1929.

63.  Ives,  K. J.  Progress in filtration.  J. Am. Water Works Assoc.,
     16(9):1225-1232,  1964.

64.  Ives,  K. J.  Theory of filtration.  Special Subject No. 7.  Proc.
     Int. Water Supply Cong. Exhib., Vienna, 1969.  London, International
     Water  Supply  Association,  1969.

65.  Ives,  K. J.  Personal communication,  September 24, 1974.

66.  Jackson, R.  The  mechanics of fluidized beds.  Part I: The stabil-
     ity of the state  of uniform fluidization.  Trans. Inst. Chem.
     Eng.,  41:13-21, 1963.

67.  Johnson, R. L., and Cleasby, J. L.  Effect of backwash on filter
     effluent quality.  ASCE J. Sanitary Eng.  Div., 92:215-228, 1966.

68.  Joslin,  J. R., and Greene, 6.  Sand filter experiments at Derby.
     Water  Pollut. Control, 69:611-622,  1970.

69.  Jottrand, R.  An  experimental study of the mechanism  of fluidiza-
     tion.   J. Appl. Chem., ^(Supplementary Issue,1):5l7-526, 1952.

70.  Jung,  H., and Savage, E. S.  Deep-bed filtration.  J. Am. Water
     Works  Assoc., 66:73-78, 1974.

71.  Kada,  H., and Hanratty, T. J.  Effects of solids on turbulence in
     a  fluid.  AIChE J., <>:624-630.  1960.

72.  Reefer,  C. E. Tertiary sewage treatment.  Public Works, 93(11):
     109-112, 1962.

73.  Kramers, I. H., Westermann, M. D.,  de Groot, J. H., and Dupont,
     F. A.  A.  The longitudinal dispersion of  liquid in a  fluidized
     bed.   Symposium on the Interaction  Between Fluids and Particles.
     London,  Institution of Chemical Engineers, pp. 114-119, 1962.

74.  Le'Clair, Brian P. Two component fluidization.  Unpublished M.A.
     Sc. thesis.  British Columbia, Canada, Library, University of
     British Columbia, 1964.

75.  Lemlich, R.,  and  Caldas, I.  Heat transfer to a liquid fluidized
     bed.  AIChE J.,  4:376-380,  1958.

76.  Leva,  Max.  Fluidization.   New York,  McGraw-Hill Book Co., Inc.,
     1959.
                                   349

-------
77.  Leva, M., Grummer, M.,  Weintraub, M.,  and Pollchik,  M.   Intro-
     duction to fluidization.   Chem.  Eng.  Prog.,  44:511-520,  1948.

78.  Leva, M., Grummer, M.,  Weintraub, M.,  and Pollchik,  M.   Fluidi-
     zation of solid nonvesicular particles.   Chem.  Eng.  Prog.,  44;
     619-626, 1948.

79.  Lewis, E. W., and Bowerman, Ernest W.  Fluidization of solid
     particles in liquids^  Chem. Eng. Prog., 48:603-609, 1952.

80.  Malik, A. M.  Air and water backwashing of granular filters.
     Unpublished M.S. thesis.   Ames,  Iowa, Library,  Iowa State Univer-
     sity of Science and Technology,  1972.

81.  McCune, L. K., and Wilhelm, R. H.  Mass and momentum transfer in
     solid-liquid system.  Ind. Eng.  Chem., 41:1124-1134, 1949.

82.  Metcalf and Eddy, Inc.  Wastewater Engineering.  New York,
     McGraw-Hill Book Co., Inc., 1972.

83.  Michaelson, A. P.  Under the solids limit at Ashton-under-Lyne.
     Water Pollut. Control, 70(5)-.533-535, 1971.

84.  Mohanka, S. S.  Theory of multilayer filtration.   ASCE J. Sanitary
     Eng. Div., 95:1079-1095,  1969.

85.  Murray, J. D.  Mathematical aspects of bubble motion in fluidized
     beds.  Chem. Eng. Prog. Symp. Ser., £2(62):71-82, 1966.

86.  Naylor, A. E., Evans, S.  C., and Dunscombe,  K.  M.  Recent develop-
     ments on the rapid sand filters  at Luton, Water Pollut.  Control,
     66(4):309-320, 1967.

87.  Nicolle, N. P.  Humus tank performance,  microstraining and  sand
     filtration.  Inst. Sewage Purifi., J. Proc., .1:19, 1955.

88.  Nicolle, N. P.  Operational aspects of control  of rapid  gravity
     sand filters.  Inst. Sewage Purif., J. Proc., 3:273-275, 1957.

89.  Ostergaard, K.  The effect of particle size  and bed  height  on the
     expansion of mixed phase  (gas-liquid)  fluidized beds.  Chem. Eng.
     Sci., .21:413-417, 1966.

90.  Ostergaard, K.  On the growth of air bubbles formed  at a single
     orifice in a waterfluidized bed.  Chem.  Eng. Sci., 21:470-472.
     1966.

91.  Othmer, D. F.  Fluidization.  New York,  Reinhold  Publishing Corpor-
     ation, 1956.

92.  Pettet, A. E. J., Collett,  B.  A., and Summers,  T. H.  Mechanical
     filtration of sewage effluents.   Part 1, Removal  of  humus.   Inst.
     Sewage Purif., J. Proc.,  4:399-410.  1949.

                                  350

-------
93.  Pettet, A. E. J., Collett, W. F., and Waddington,  J.  I.   Mechan-
     ical filtration of sewage effluents.   III.  Removal of humus:
     Further experiments at Luton.  Inst.  Sewage Purif., J. Proc.,
     4:195-202, 1951.

94.  Pettet, A. E., Collett, W. F.,  and Waddington,  J.  I.   Rapid filtra-
     tion of sewage effluents through sand and anthracite. Sewage   Ind.
     Wastes, 24(7):835-843, 1952.

95.  Pinchbeck, P. H., and Popper, F.  Critical  and  terminal  velocities
     in fluidization.  Chem. Eng. Sci., 6.:57-64, 1956.

96.  Pruden, B. B.  Particle size segregation in particulately fluid-
     ized beds.  Unpublished M. A. Sc. thesis.  British Columbia,
     Canada, Library, University of British Columbia,  1964.

97.  Pruden, B. B., and Epstein, N.   Stratification  by size in particu-
     late fluidization and in hindered settling.  Chem. Eng.  Sci.,
     ^9:696-700, 1964.

98.  Reuter, H.  On the nature of bubbles  in gas and liquid fluidized
     beds.  Chem. Eng. Prog., Symp.  Ser., 6i2(62):92-99,  1966.

99.  Rice, G. A.  Direct filtration of secondary effluent, backwashing
     and performance.  Unpublished M.S. thesis.   Ames,  Iowa,  Library,
     Iowa State University of Science and  Technology,  1973.

100. Richardson, J. F., and Zaki, W. N.  Sedimentation and fluidization:
     Part I.  Trans. Inst. Chem. Eng., .32:35-53, 1954.

101. Rigby, G. R., Blockland, G. P.  V., Park, W. H., and Capes, C.  E.
     Properties of bubbles in three phase  fluidized  beds as measured by
     an electro-resistivity probe.  Chem.  Eng. Sci., 25:1729-1741,  1970.

102. Ripley, P. G., and Lamb, G. L.   Filtration of effluent from a
     biological-chemical system.  Water Sewage Works,  120(2); 66-69,
     1973.

103. Robeck, G. G., and Kreissl, J.  F.  Multi-media  filtration: princi-
     ples and pilot experiments.  Transactions of 17th Annual Conference
     of Sanitary Engineering.  Univ. Kans. Public Bull. Eng.  Arch.,
     5_7:lO-22, 1967.

104. Romero, J. B., and Johanson, L. N.  Factors affecting fluidized
     bed quality.  Chem. Eng. Prog., Symp. Ser., 58(38):28-37, 1962.

105. Rowe, P. N.  Drag forces in a hydraulic model of a fluidized bed.
     Part II.  Trans. Inst. Chem. Eng., 39.: 175-180,  1961.

106.  Rowe, P. N., and Henwood, G. A.  Drag forces in a hydraulic
      model of a fluidized bed.  Part I.  Trans. Inst. Chem. Eng., 39;
      43-54, 1961.
                                   351

-------
 107.  Ruckenstein, E.  Homogeneous f lutdizatlon.  Ind. Eng. Chem. Fund.,
      3:260-268, 1964.

 108.  Savage, E. S., Personal communication, February 19, 1975, describ-
      ing media and backwash provisions used by Pintsch Bamag (Germany)
      and Dravo (U.S.) .

 109.  Sejkora, G. D.  Effect  of Media  Intermixing on Dual Media Filtra-
      tion.   Unpublished M. S. thesis.   Ames,  Iowa,  Library,  Iowa State
      University of Science and Technology,  1974.

110. Shannon, P. T.  Fluid dynamics of gas fluidized batch systems.
     Microfilm copy.  Unpublished Ph.D. thesis.  Chicago, Illinois,
     Library, Illinois Institute of Technology, 1961.

111. Shi reman, C.  Filtration boosts tertiary treatment.   Water Wastes
        ., £(4):34-37, 1972.
112. Siusnonds, M. A.  Public water supplies of Queensland, Australia,  J.
     Am. Waterworks Assoc., 55_: 1044 -1080, 1963.

113. Simpson, H. C., and Rodger, B. W.  The fluidization of light solids
     by gases under pressure and heavy solids by water.  Chem.  Eng.,
     Sci., 16_:153-180, 1961.

114. Singer, P. C., and Stumm, W.  The solubility of ferrous iron in
     carbonate-bearing waters.  J. Am. Water Works Assoc., 62; 198-202.
     1970.

115. Slis, P. L., and Willemse, T. W., and Kramers, H.  The response of
     the level of a liquid fluidized bed to a sudden change in the
     fluidizing velocity, Appl. Sci. Res., 8:209-218, 1959.

116. Sosewitz, B., and Bacon, V. W.  Chicago's first tertiary treatment
     plant.  Water Wastes Eng., 5_:52-55, 1968.

117. Standard Methods for the Examination of Water and Wastewater-
     13th ed. New York, American Public Health Association, 1971.

118. Stangl, E. W.  Air and water backwash of wastewater filters.  Un-
     published M.S. thesis.  Ames, Iowa, Library, Iowa State University
     of Science and Technology, 1973.

119. Steward, P. S., and Davidson, J. F.  Three-phase fluidization:
     Water, particles and air.  Chem. Eng. Sci., 12:319-322, 1964.

120. S too key, L. L.  Ferrozine - A new spectrophotometric reagent for
     iron.  Anal. Chem., 42:779-781, 1970.

121. Streander, B.  Mechanical filtration of sewage.  Water Works
     Sewerage, 82(7) : 25 2- 257, 1935.
                                   352

-------
122. Streander, B.  Sewage filtration with silica sand filters  I.  Water
     Works Sewerage, 87(8)-.351-357,  1940.

123. Streander, B.  Sewage filtration with silica sand filters  II.
     Waterworks Sewarage, 87(9)-.422-429,  1940.

124. Streander, B.  Sewage filtration with silica sand filters  III.
     Waterworks Sewerage, 87(10)-.493-495, 1940.

125. Taylor, G. I.  Statistical theory of  turbulence,  Parts  I-IV.
     Proc. R. Soc.,London, Ser. A.,  l5JL:421-478,  1935.

126. Taylor, G. I.  The spectrum of  turbulence.   Proc. R.  Soc.  London,
     Ser. A, 164:476-490, 1938.

127. Tchobanoglous,  G.  Filtration techniques in  tertiary  treatment.
     J. Water Poll.  Control Fed.,  42:604-623, 1970.

128. Tchobanoglous,  G., and Eliassen, R.  Filtration of treated sewage
     effluent.  ASCE J. Sanitary Eng. Div., 96:243-265, 1970.

129. Tebbutt, T. H.  Y.  An investigation into tertiary treatment by
     rapid filtration.  Water Res.,   J. Intern. Assoc. Water Pollution
     Research, 5_:81-92, 1971.

130. Tesarik, I.  (Discussion of)  Theory of water filtration. ASCE J.
     Sanitary Eng. Div., 90:67-69, 1965.

131. Truesdale, G. A., and Birkbeck, A. E.  Tertiary treatment  of
     activated-sludge effluent.  Water Pollut. Control, 67(5):483-494,
     1968.

132. Ulug, S. E.  The backwaahing of rapid sand filters.  Unpublished
     thesis for the Diploma of membership of Imperial College  of Science
     and Technology.  London, United Kingdom, Library, Imperial College
     of Science and Technology, University of London, 1967.

133. United States Environmental Protection Agency.  Process Design
     Manual for Suspended Solids Removal.   U.S. EPA Technology  Transfer,
     1975.

134. Van Heerden, C., Novel, A. P. P., and Van Krevelen, D.  W.   Studies
     on fluidization - The critical mass velocity.  Chem.  Eng.  Sci.,
     l.:37-49, 1951.

135. Volpicelli,  G., Massimilla, L. and Zenz, F.  A.  Non-homogeneities in
     solid-liquid fluidization.  Chem. Eng. Frog., Symp. Ser.,  62(67):
     42-50, 1966.

136. Vosloo, P. B. B.  Some experiments on rapid sand filtration of
     sewage works effluents without coagulation.   Inst. Sewage Purif., J.
     Proc., 1:204-209, 1947.
                                   353

-------
137. Walker, J. D., Aurora, Illinois.  Dual media tertiary filters
     design parameters.  Private communication with Walker Process,
     August 1, 1972.

!38. wen, C. Y., and Yu, Y. H.  Mechanics of fluidization.  Chem.  Eng.
     Prog., Symp. Ser., .62(62) : 100-111, 1966.

139. Whitmore, R. L.  The sedimentation of suspensions of spheres.  Br.
     J. Appl. Phys., 6:239-245, 1955.

140. Whitmore, R. L.  The relationship of the viscosity to the settling
     rate of slurries.  J. Inst. Fuel, 30:328-342, 1957.

141. Wilhelm, R. H., and Kwauk, M.   Fluidization of solid particles.
     Chem. Eng. Prog., 44:201-218,  1948.

142. Wood, R., Smith, W. S., and Murray,  J.  K.   An investigation into
     upward flow filtration.  Water Pollut.  Control,  6_7(4) :42l-428,
     1968.

143. Woods, C. F.  Expansion and intermixing of garnet and silica  sand
     during backwashing of granular filters.  Unpublished M.S.  thesis.
     Ames, Iowa, Library, Iowa State University of Science and Tech-
     nology, 1973.

144. Zabrodsky, S.  S.  Hydrodynamics and  Heat Transfer in Fluidized
     Beds.  Cambridge, Massachusetts, The M.I.T. Press,  1966.

145. Zenz, F. A. and Othmer, D. F.   Fluidization and  Flufil Particle
     Systems.  New  York, Rheinhold  Publishing Corporation, 1960.
                                  354

-------
      Sieve analyses of uniform media.
U>

Designation
F(-4+7)


F(-7+8)


F(-8+10)


F( -10+12)


A(-12+14)


A( -14+16)


A(-16+18)


A(-18+20)


AC -20+25)
Coals
U.S.
sieve
4
7
Pan
7
8
Pan
8
10
Pan
10
12
Pan
12
14
Pan
14
16
Pan
16
18
Pan
18
20
Pan
20
14
Coals (continued)
retained
3.4
79.7
16.9
3.3
85.0
11.7
4.6
81.0
14.4
4.8
84.2
11.0
4.7
80.5
14.8
2.2
83.6
14.2
7.0
77.8
15.2
3.8
77.8
18.4
10.2
Designation U.S. %
sieve retained
A(-25+30) 25
30
Pan
A( -30+35) 30
35
Pan
A(-35+40) 35
40
Pan
A(-40+45) 40
45
Pan
14
Silica sands

A(-10+12) 10
12
Pan
A(-12+14) 12
14
Pan
A(-14+16) 14
16
Pan
A(-16+18) 16
4.9
84.1
11.0
3.2
83.3
13.5
8.8
81.7
9.5
4.2
83.6
12.2



0.1
83.7
16.2
9.2
80.1
10.7
5.9
70.1
24.0
2.7
14
Silica sands (continued)
Designation U.S.
sieve
A(-18+20) 18
20
Pan
A(-20+25) 20
25
Pan
A(-25+30) 25
30
Pan
A (-30+35) 30
35
Pan
A(-35+40) 35
40
Pan
A (-40+45) 40
45
Pan

Garnet sands

A(-14+16) 14
16
18
Pan
retained
4.3
87.0
8.7
0.6
74.9
24.5
2.1
77.6
20.3
1.3
70.9
26.8
3.6
72.6
23.8
6.2
74.8
19.0
143
i^Tn^

0.24
87.72
12.00
0.04

-------
    Sieve analysis of uniform media (continued).
tn


Designation



A(-25+30)



A(-50460)



14
Coals
U.S.
sieve
25
Pan
25
30
35
Pan
50
60
70
Pan

1A

Coals'""' (continued)
%
retained
78.3
11.5
0.11
88.91
10.46
0.53
0.14
81.81
16.56
1.49
Designation U.S.
sieve
18
Pan
A(-20+25) 20
25
30
Pan
A(-30+35) 30
35
40
Pan
%
retained
74.5
22.8
0.22
78.40
21.33
0.05
0.54
88.76
10.58
0.12
14

Silica sands (continued)
Designation U.S.
sieve


A(-35-»40) 35
40
45
Pan
A (-40445) 40
45
50
Pan
%
retained


0.42
85.17
14.33
0.08
0.70
96.12
3.15
0.03

-------
                                   TECHNICAL REPORT DATA
                            (Please read Instructions on the reverse before completing)
1. REPORT NO.
   EPA-600/2-77-016
                             2.
             3. RECIPIENT'S ACCESSION-NO.
4. TITLE AND SUBTITLE

   BACKWASH OF  GRANULAR FILTERS USED IN WASTEWATER
   FILTRATION
             5. REPORT DATE
              April 1977  (Issuing Date)
             6. PERFORMING ORGANIZATION CODE
7. AUTHOR(S)


   J..L. Cleasby and E.  R. Baumann
             8. PERFORMING ORGANIZATION REPORT NO.
9. PERFORMING ORGANIZATION NAME AND ADDRESS
   Iowa State University
   Ames, Iowa 50011
             10. PROGRAM ELEMENT NO.

              1BB043.ROAP  21-ASO.Task 15
             11. CONTRACT/GRANT NO.
                                                            R802140
 12. SPONSORING AGENCY NAME AND ADDRESS
 Municipal Environmental Research Laboratory--Cin.,OH
 Office of Research and Development
 U.S. Environmental Protection Agency
 Cincinnati, Ohio  45268
             13. TYPE OF REPORT AND PERIOD COVERED
              Final,  9/1/71-5/31/76	
             14. SPONSORING AGENCY CODE
              EPA/600/14
15. SUPPLEMENTARY NOTES
16. ABSTRACT
 The use of deep  granular filters in waste  treatment is of growing importance.   The
 key to long-term operating success of such filters is proper bed design  and adequate
 bed cleaning during backwashing.  A number of questions related to adequate backwash-
 ing of granular  filters are investigated which lead to the following  conclusions:
 Cleaning granular filters by water backwash alone to fluidize the filter bed is
 inherently a weak cleaning method because  particle collisions do not  occur  in  a
 fluidized bed  and thus abrasion between the filter grains is negligible.
 Due to the inherent weakness of water backwashing cited above, auxiliary means of
 improving filter bed cleaning are essential for wastewater filters.   Three  auxiliary
 methods were compared in a wastewater pilot filtration study.  The most  effective
 backwash was provided by air scour and water backwash simultaneously  at  subfluidiza-
 tion velocities.   The other two methods, surface and subsurface wash  auxiliary or
 air scour prior  to water fluidization wash were about comparable in effectiveness.
 The performance  of coarse sand, dual-, and triple-media filters was compared,  and
 the backwashing  routines appropriate for each media are discussed.  A number of
 investigations concerning the design and backwashing of dual media filters  are
 presented.
17.
                                KEY WORDS AND DOCUMENT ANALYSIS
                  DESCRIPTORS
                                              b.lDENTIFIERS/OPEN ENDED TERMS
                             COS AT I Field/Group
 Filtration  , Sewage Filtration ,
 Sewage Treatment,  Waste Treatment
 Water Pollution,  Backwashing*,
 Clarification
 Filter Backwash, Sus-
 pended Solids Removal,
 Filter Media, Filter
 Cleaning, Filter Design
      13B
18. DISTRIBUTION STATEMENT


 RELEASE TO PUBLIC
19. SECURITY CLASS (This Report)
 UNCLASSIFIED
ZO. SECURITY CLASS (This page)
 UNCLASSIFIED
21. NO. OF PAGES

      381
                                                                         22. PRICE
EPA Form 2220-1 (9-73)
                                            357
                                                                     -1977-757-056/5573 Region No. 5-11

-------