U.S. Department
of Transportation
Federal Aviation
Administration
Nitric Oxide
Measurement  Study
Office of Environment
and Energy
Washington, D.C. 20591
Probe Methods
Volume II
Report Numbers:
FAA-EE-80-29
USAF ESL TR-80-13
NASA CR-159862
USN NAPC-PE-38C
EPA-460/3-80-014
MAY 1980
M.B. Colket, III
M.F. Zabielski
L.J. Chiappetta
L.G. Dodge
R.N. Guile
D.J. Seery

-------
This document is disseminated under the joint sponsorship
of the Federal Aviation Administration, U.S. Air Force,
U.S. Navy, National Aeronautics and Space Administration,
and the Environmental Protection Agency in the interest of
Information exchange.  The United States Government assumes
no liability for the contents or use thereof.

-------
                                                              Technical Report Documentation Page
  FAA-EE-80-29
  4. Title and Subtit
                                2.  Government Accession No.
  Nitric Oxide Measurement Study:  Probe Methods -
  Volume  II
  7. Author's) M_ B.  Colket, III, M. F.  Zabielski,
 J-.  J.  Chiappetta,  L.  G.  Dodge, R.  N.  Guile, D. J. Seery
  9. Performing Orgomiotion Nome and Address
  United Technologies Research Center
  Silver Lane
  East Hartford,  CT   06108
  12.  Sponsoring Agency Nome ond Address
  U.S. Department  of Transportation
  Federal Aviation Administration
  Office of Environment and Energy
 .Washington.  DC    20591	
                                                             3. Recipient's Cotalog No.
                                                             5. Report Date
                                                              March 31,  J98Q
                            6.  Performing Organization Code
                                                             8.  Performing Organization Report No.
                                                            R79-994150-2
                            10. Work Unit No. (TRAIS)
                            11. Contract or Grant No.
                            DOT FA77WA-4081
                                                            13. Type of Report and Period Covered
                            14. Sponsoring Agency Code
  15. Supplementary Notes
                    Funding for this  study was provided by an  Interagency Committee.
  Contributing  agencies and report nos.  are:  DOT-FAA (FAA-EE-80-29); USAF (ESL TR-80-
  13); NASA  (CR-159862); USN  (NAPC-PE-38C);  and EPA  (EPA-460/3-80-014).
  16 Abstract                         ~
  Experimental  facilities used in  studying the performance  of  probes and sampling
  systems for measuring NO are described.   A critical review of  the literature on probe
  measurements  of  NOX is given with  emphasis on reported  results indicating that probes
  may perturb the  total concentration of  NOX in a flame.  Also,  sample line and
  chemiluminesct-nt analyzer phenomena are  reviewed.  A model of  probe aerodynamics
  including heat  transfer is presented.   Kinetics of NO loss are examined and quenching
  criteria for  measuring nitric oxide in  flames are given.  Sampling probes are de-
  scribed that  were designed to preserve  NO and are suitable for measurements on small
  and large combustors.  Probes were designed to cool the gases  both convectively and
  aerodynamically.   Performance of these  probes is compared with model predictions.
  Concentrations  of nitric oxide were measured using several probes for each of three
  flame environments.  The values  measured with each probe  are compared and related
  to seed levels  of NO.  In addition, concentration profiles required to compare probe
  measurements  with optical measurements  are provided.

  The Nitric Oxide Measurement Study is in three volumes:
  Optical Calibration - Volume I;
  Probe Methods - Volume II;
  Comparison of Optical and Probe  Methods - Volume III.
 17. Key Words
  Nitric oxide,  probe sampling,  chemilumin
  escent analysis,  aerodynamic  analysis.
             18. Distribution Statement
              Document is available  to public through
              the National Technical Information
              Service, Springfield,  VA   22161
 19.  Security Clossif. (of thil report
  Unclassified
20. Security Classif. (of this page)
Unclassified
21. No. of Poges

  112
22. Pr
Form DOT F 1700.7  (8-72)
                              Reproduction of completed poge outhorized

-------
                             ACKNOWLEDGMENTS

     This contract was administered by the Federal Aviation Administration.
Funding for this work was provided by an Interagency Committee representing
the Federal Aviation Administration (FAA), Air Force, Navy, National
Aeronautics and Space Administration (NASA), and the Environmental Protection
Agency (EPA).

     The assistance of Mr. D. L. Kocum, Mr. R. P- Smus, Mr. D. D. Santos and
Mr. R. L. Poitras during the experimental portions of this study is gratefully
acknowledged.  The authors also would like to acknowledge the contributions of
the following UTRC staff:  Mrs. B. B. Johnson and Mr. C. Foley for data reduction
and report preparation; and Messrs. P. N. Cheimets, M. E. Maziolek, W. T. Knose,
and M. Cwikla for facilities support.

-------
                                   ABSTRACT
     Experimental facilities used in studying the performance of probes
and sampling systems for measuring NO are described.  A critical review of the
literature on probe measurements of NO  is given with emphasis on reported
results indicating that probes may perturb the total concentration of N0y in
a flame.  Also, sample line and chemiluminescent analyzer phenomena are reviewed.
A model of probe aerodynamics including heat transfer is presented.  Kinetics
of NO loss are examined and quenching criteria for measuring nitric oxide in
flames are given.  Sampling probes are described that were designed to preserve
NO and are suitable for measurements on small and large combustors.  Probes
were designed to cool the gases both convectively and aerodynamically.  Perfor-
mance of these probes is compared with model predictions.  Concentrations of
nitric oxide were measured using several probes for each of three flame envi-
ronments.   The values measured with each probe are compared and related to seed
levels of NO.  In addition, concentration profiles required to compare probe
measurements with optical measurements are provided.
                                        ii

-------
                               TABLE OF CONTENTS
ACKNOWLEDGEMENTS   	   i

ABSTRACT	   ii

TABLE OF CONTENTS	   iii

LIST OF FIGURES    	   v

LIST OF TABLES	   vi

  I.  INTRODUCTION	   1-1

 II.  EXPERIMENTAL FACILITIES  	   II-l

      A.  General	   II-l
      B.  Flat Flame Burner	   II-l
      C.  Test Section for Large Scale Burners	   II-3

          1.  IFRF Burner	   II-3
          2.  FT12 Burner Can	   II-8
          3.  Temperature Measurements	   II-8

      D.  Sampling Systems  	   11-12

          1.  Scott Exhaust Analyzer    	   11-12
              a.  Pumping Requirements  	   11-13
          2.  TECO Analyzer	   11-13

      E.  Mass Flow Measurement	   11-14

III. DESIGN OF GAS SAMPLING PROBE	   III-l

     A.   Losses of Nitric Oxide in Sampling System  	   III-l

         1.   N0/N02 Interconversion 	   III-2
             a.   The Bodenstein Reaction	   III-4
         2.   NO  Reduction in Sampling Probe	   III-8
               X
         3.   Losses in Sampling Line	   111-13
         4.   Response of Chemiluminescence Analyzer 	   111-14
         5.   NO  Converter	   111-16
         6.   Summary	   III-17
                                        iii

-------
                           TABLE  OF  CONTENTS  (Cont'd)
    B.  Quenching  in Gas  Sampling  Probe  	   111-17

        1.  Kinetics of NO  Decomposition	111-20
        2.  Description of  Computer  Program  for  Probe  Analysis.  .   111-24
            a.   Sudden Expansion Losses  	   111-26
            b.   Aerodynamic Quench	111-27

    C.  Design  of  Probes	   111-30

        1.  Probes  for Combustor Measurements  	   111-30
        2.  Probes  for the  Flat Flame Burner	111-35
            a.  Microprobe  	   111-42

 IV. EXPERIMENTAL RESULTS   	   IV-1

    A.  Flat Flame  Burner	   IV-1
        1.  Uncooled, Stainless Steel Probe  	   IV-12

    B.  IFRF Burner	IV-12

    C.  FT12 Measurements	   IV-20

    D.  Experimental Verification  of Probe Model   	   IV-25

        1.  Pressure Profiles  for  the Reference  Probe  	   IV-25
        2.  Mass Flow Measurements	IV-28
            a.   Macroprobe	   IV-28
            b.   Miniprobe   	   IV-28
            c.   Microprobe	   IV-30
        3.  Discussion	   IV-30

 V. RESULTS AND  DISCUSSION	V-l

VI. CONCLUSIONS	VI-1

    REFERENCES	R-l

-------
                                LIST OF FIGURES
Fig. No.                    Title
 II-l.              Top View of Flat Flame Burner and Assembly	II-2
 II-2.              Atmospheric Pressure Combustion Facility	II-5
 II-3.              Swirl Burner Assembly 	  II-6
 11-4.              FT12 Assembly	II-9
 II-5.              Pt-Pt/13% Rh Aspirated Thermocouple 	  11-11
 II-6.              Experimental Set-up for Measurements of Mass Flow .  11-15
 II-7.              Indicated Reading at Constant Mass Flow and
                      Varying Operating Pressure  	  11-16
 II-8.              Metered Mass Flow Rates vs. Flow Meter Reading.  . .  11-17

III-l.              Model for Calculating Sudden Expansion Loss .... 111-28
III-2.              Drawings of Macroprobes 	 111-31
III-3.              Reference Probe 	 111-32
III-4.              Tip of Reference Probe	111-33
III-5.              Calculated Temperature and Pressure Profiles
                      for Reference Probe 	 111-34
III-6.              Calculated Cooling Curves for Macroprobes  	 111-36
III-7-              Drawing of Miniprobe  	 111-38
III-8.              Stainless Steel Tipped Miniprobe  	 111-39
III-9.              Calculated Cooling Curves for Miniprobe 	 111-40
111-10.             Calculated Cooling Curves for Miniprobe at
                      Varying Back Pressure 	 111-41

 IV-1.              Horizontal Temperature Profile over CH,/02/N2
                       Flat Flame	IV-2
 IV-2.              Vertical Temperature Profile over CH4/02/N2
                      Flat Flame  	IV-3
 IV-3.              Normalized Nitric Oxide Profiles over CH4/02/N2/NO
                      Flat Flame	IV-6
 IV-4.              Vertical Profiles of Nitric Oxide over Flat Flame
                      Burner	IV-7
 IV-5.              Nitric Oxide Measured vs.  Nitric Oxide Seed ....  IV-10
 IV-6.              Temperature Profile Across IFRF Combustor  	  IV-15
 IV-7-              Normalized Nitric Oxide Profiles Across IFRF
                      Combustor	IV-18
 IV-8.              Temperature Profiles Downstream of FT12 Combustor .   IV-21
 IV-9.              Normalized Nitric Oxide Profiles Across Optical
                      Axis for FT12 Combustor	IV-24
 IV-10.             Profiles of Static Pressure for the Reference
                      Probe   	IV-26
 IV-11.             Relative Mass Flow vs.  Back Pressure for Several
                      Probes	IV-29

-------
                                 LIST OF TABLES
Table No.
        Title
 II-A.
 II-B.
 II-C.

III-A.
III-B.
III-C.
III-D.

 IV-A.

 IV-B.

 IV-C.
 IV-D.

 IV-E.
 IV-F.
Operating Conditions for the Flat Flame Burner. . .    II-4
Operating Conditions for the IFRF Burner	    II-7
Operating Conditions for the FT12 Combustor ....    11-10

Flow Conditions in Non-Ideal Sampling System.  . . .   III-6
Comparison of Calculated and Measured NO and NOo. .   III-9
Reaction Mechanism for NO Decomposition 	   111-21
Estimated Fractions of NO Decomposition 	   111-23

Mole Percent of Stable Species for Flat Flame
  Burner	    IV-4
Measured Concentration of NO(ppm) Using Uncooled,
  Stainless Steel Probe over Flat Flame Burner. . .    IV-13
Mole Percent of Stable Species for IFRF Burner. . .    IV-16
Comparison of Nitric Oxide Measurements using
  the Reference Probe - IFRF Burner	     IV-19
Mole Percent of Stable Species for FT12 Combustor .     IV-22
Comparison of Nitric Oxide Measurements using
  the Reference Probe - FT12 Combustor	     IV-25
                                       vi

-------
                                I.   INTRODUCTION
     Since Johnston  (1971) and Crutzen  (1970,  1972)  independently  suggested
that the  injection of nitric  oxide  (NO)  into  the  upper  atmosphere  could  signif-
icantly diminish the ozone (Oo)  concentration, an accurate  knowledge  of  the
amount of NO emitted by jet aircraft has  been  a serious  concern  to those
involved  in environmental studies.  This  concern  intensified when  McGregor,
Seiber, and Few (1972) reported  that NO  concentration measured by  ultraviolet
resonant  spectroscopy were factors  of  1.5 to  5.0  larger  than those measured  by
extractive probe sampling with subsequent chemiluminescent  analysis.   These
initial measurements were made on a YJ93-GE-3  engine as  part of  the Climatic
Impact Assessment Program (CIAP) which  was  one of four  studies (CIAP,  NAS,
COMASA, COVOS  (see References))  commissioned  to determine the possible environ-
mental consequences  of high altitude aircraft  operation, especially supersonic
aircraft.  After those studies were initiated, economic  factors  strongly
favored the production and operation of  subsonic  aircraft.  Nevertheless,  since
the  subsonic aircraft fleet is large and  does  operate as high as the  lower
stratosphere,  interest in the causes of  the discrepancies between  the  two NO
measurement methods  continued.   Few, Bryson, McGregor,  and  Davis (1975, 1976,
1977) reported a second set of measurements on an experimental jet  combustor
(AVCO-Lycoming) where the spectroscopically determined NO concentrations were
factors of 3.5 to 6.0 higher  than those  determined by the probe method.  In
this set  of measurements, optical data were obtained not only across  the
exhaust plume  but also in the sample line connecting the probe with the chemilum-
inescent  analyzer.   The sample line optical data  seemed  to  agree with  the
chemiluminescent analyzer data;  hence,  it was  suggested  that the discrepancies
were due  to phenomena occurring  in  the  probe.  These results stimulated a third
set  of measurements  involving ultraviolet spectroscopy  (Few et al,  1976a,
1976b), infrared gas correlation spectroscopy  (D.  Gryvnak,  1976a,  1976b) and
probe sampling on an Allison  T-56 combustor.   The measured  ratios  of  the
ultraviolet to the probe values  typically ranged  between 1.5 and 1.9 depending
on the data reduction procedure.  The  ratios of the  infrared to  the probe
values varied  between 1.1 to  1.5 also  depending on the method of data  reduction.
In addition to these engine and  combustor data, evidence supportive of the
accuracy of the ultraviolet spectroscopic method,  i.e.,  calibration data and
model predictions, was presented by McGregor,  Few, and Litton (1973);  Davis,
Few, McGregor  and Classman (1976);  and Davis,  McGregor,  and Few  (1976).
Nevertheless,  it was still not possible  to make a judgment  on the  relative
accuracy of the spectroscopic and probe methods.   The most  significant reasons
for  this were:  the  complexity of the  spectroscopic  theory  and computer model
required to infer concentration  from optical transmission;  the inadequate
treatment of probe use; and the  incomplete exhaust temperature and  pressure
data which are necessary for  a valid comparison of the methods.  Recently,
Oliver et al (1977,  1978) as  part of the  High  Altitude Pollution Program has
ranked these discrepancies as a major  and  a continuing source of uncertainty in
atmospheric model predictions.
                                       1-1

-------
     The purpose of this investigation was to identify and determine  the
magnitude of the systematic errors associated with both the optical and probe
sampling techniques for measuring NO.  To accomplish this, the study  was
divided into three parts.  The first was devoted to calibrating the ultraviolet
and infrared spectroscopic methods.  This entailed developing procedures which
could provide known concentrations of NO over a wide range of temperatures  and
pressures, and also reviewing and correcting the ultraviolet spectroscopic
theory used in the engine and combustor measurements cited above.  The second
part of this study was focused on sample extraction, transfer, and analysis  of
chemiluminescent instrumentation.  The sampling methods were used on  three
successively more complicated combustion systems starting with a flat flame
burner and culminating with a jet combustor.  The results are presented in TASK
II Report: Probe Methods.  In the third part of this study, optical measurements
were made on the same three combustion systems operated at the same conditions
used for the probe measurements.  The results of the optical and probe measure-
ments were compared and are given in TASK III Report:  Comparison of Optical
and Probe Methods.

     This report, i.e., TASK II, is devoted to the important processes in the
extractive sampling and measurement of NO from combustion streams.   Described
are a flat flame burner, swirl combustor, and jet combustor along with their
support facilities and operating parameters.  Information on the temperature,
analytical, and mass flow instrumentation is given.  Problems associated with
loss of nitric oxide in a sampling system are reviewed and the results of
previous investigations are analyzed.  Several different probes were used for
sampling the flame gases, and their designs were selected using a computer
program describing the principal aerodynamic and heat transfer processes
encountered in a probe.  Direct experimental measurements of the fluid mechanics
within probes is presented verifying this model.  In addition,  temperature and
concentration profiles necessary to compare optical and probe measurements are
provided.   These results are summarized and discussed and major conclusions are
given.
                                       1-2

-------
                          II.   EXPERIMENTAL  FACILITIES
II. A.  General
     A principal objective  of  this  study  is  to  identify  the  relative  merits
of probe and UV optical measurements  of NO in the  exhaust  of aircraft engines.
To reach this goal,  three combustion  systems  of varying  degrees  of  complexity
were examined.  These  systems  were:
      1.  CH^/C>2/N2  flame  over  a  flat  flame  burner:
            =  0.8,  1.0,  1.2,  P  =  1  atm,  mTOTAL  s  2.75  g/sec

      2.  C-jHg/Air  flame  in  a  swirl  burner:

         4,  = 0.8,  1.0, 1.2, Swirl = 0.63 and  1.25,  P = 1  atm, m-p^j^ s  71g/sec
     3.  Jet A/Air  flame  in  a modified  FT12  combustor:  Idle,  Cruise,
         and Maximum  Continuous, P  =  1  atm,
     Physical details  and  operation  of  the  flat  flame  burner have been described
previously  (Dodge, et  al . ,  1979);  consequently,  only a brief overview will be
presented here.   For the  other  two  flames,  each  burner assembly  could be
installed separately into  a single  combustor  housing with  the  associated  fuel
lines  and flow  controls modified  accordingly.  This  facility and the burner
assemblies  are  described  in detail  in this  chapter.  In addition, techniques
for  temperature measurements  with  corrections, sample  gas  transfer  and analysis,
and  mass flow measurements  are  reviewed.  Details  of sample probe construction
will be discussed in the  following  chapter  since their designs were defined by
model  predictions.
II. B. Flat Flame Burner

      The  flat  flame burner  is made  of  sintered  copper  and has  two zones:   the
main  zone  (containing  the main  flame seeded with  nitric  oxide) with dimensions
of 17.5 x  9.2  cm or 161 cm   and  the  (unseeded)  buffer  zone with  an area  of  76
cm^.  A methane flame  was burned  above  the buffer flame  to provide a hot  zone
in the wings of the flame.   The  burner  was enclosed by a stainless steel
shroud/chimney with optical  ports to separate windows  (quartz  or salt)  from the
flame.  The ports were purged with  nitrogen at  room temperature  to reduce  the
local nitric oxide concentrations within  these  ports.  A top view of the  burner
is shown  in Figure II-l.  Temperatures  were measured using a butt-welded,  Ir/60%
Ir-40% Rh  thermocouple coated with  a mixture of Yttrium  and Berylium oxides.

* The stoichiometry, $, is defined  to be  (f/a)/(f/a)   ^ where  4, = ! when  the
  fuel (f) and air (a) are at the stoichiometric  ratio.
                                      II-l

-------
                                  TOP VIEW OF FLAT FLAME BURNER AND ASSEMBLY
I
r ->
                                 BUFFER ZONE-
                                                                   • S.S. SHROUD

INERT
GAS PURGE
II
1
1

MAIN GAS FLOW — ^
(SEEDED W/NO)







X








X


\

	 y 	



x


/
























>




QUARTZ WINDOWS-^
i|A
i
i

v.
^^ DEAD SPACE
                                                                                       SCALE
   to
   I
   o
5 cm

-------
The diameter of the bead and coating was approximately 90 microns  (0.0035
inches).  Gases were individually tnetered using  critical flow orifices.
Separate mixes of gases (N2, C>2, CH^, NO, H2, Ar) were blended  for  the
main and buffer flows.  Details of  these facilities are provided  in the Task  I
Report  (Dodge, et al, 1979).  The flames examined in  this program  are  listed  in
Table II-A.

II. C. Test Section for Large Scale Burners

     The combustor test section used for the  swirl burner and FT12  combustor
is shown schematically in Fig. II-2.  It consists of  a water-cooled, double-
walled  chamber 50 cm in diameter (i.d.) and  150  cm long.  Four  (4)  rows of
eight (8) viewing ports are provided in the  combustor section at 90° intervals.
It was  constructed at UTRC specifically to  investigate flame phenomena with
various optical and probing techniques.  The  two burner systems were designed
to fit  inside the burner housing.   One is a  swirl burner and is a  scaled down
verison of the burners designed at  the International  Flame Research Foundation
(IFRF).  The second is a modified FT12 burner and shroud.  The  optical axis
used for subsequent optical measurements was  the center of the  third window
from the far right in Fig. II-2.  All probes  (sample  and thermocouple) were
designed to translate across this axis.

II. C.  1  IFRF Burner

     The IFRF burner assembly, as shown in Fig.  II-3, is a model of  those burners
described by Beer and Chigier (1972) and consists of  a central  fuel nozzle and
an annular air supply.  A movable vane block  arrangement provides variable air
swirl intensity from a swirl number of 0 to  2.5; in this case,  the  swirl number
is defined as the ratio of the tangential to  the axial momentum divided by the
radius  of the exit quarl.  The swirl number  was  calculated using the appropriate
equations in the text by Beer and Chigier (1972) and, as they demonstrate,
experimental and theoretical values agree fairly well for this  type of burner.
An axially adjustable, 19 mm diameter, fuel  feed tube can be equipped with
various pressure atomizing of air-assisted  fuel  spray nozzles.

     This swirl burner has been tested previously during internal  programs at
UTRC and has been used recently to  study the  combustion of a coal/oil  slurry
(Vranos, et al, 1979).  In the present program,  gaseous propane was used for
fuel and a nozzle was constructed to inject  the  fuel  radially into  the swirling
air flow.  Six stable operating conditions were  selected for these  tests and
provide three stoichiometries and two swirls.  Input  conditions are  listed in
Table II-B.  The optical axis and the probe  tips were located 87.5  cm
downstream of the quarl exit.

     The two swirl levels used in these tests (0.63 and 1.25) were  selected by
performing a series of tests on flame stability.  At  lower swirl numbers, the
propane flame was relatively long and unstable and was not considered  to be
suitable for this series of tests.  Beer and  Chigier  argued that below a swirl
number  of 0.6 axial pressure gradients are  insufficient to cause  internal

                                       II-3

-------
                             TABLE 11-A
           OPERATING CONDITIONS FOR THE FLAT FLAME BURNER*
Test Condition               Bl               B2               B3






;N2 (g/sec)                2'15             2-20             2'07




m 02 (g/sec)                0.512            0.466            0.494





m CH4 (g/sec)               0-103            0.116            0.149




T inlec (K)                 285              285              285





P (psia)                   14.7             14.7             14.7





*                           0.8              1.0              1.2
 Without Seed
                                 II-4

-------
                                     ATMOSPHERIC PRESSURE COMBUSTION FACILITY
                                                                                                •TRANSITION DUCT
    AIR
                 SWIRL BURNER
• 4 ROWS OF 8 QUARTZ WINDOWS
WATER-COOLED, DOUBLE-WALLED. CONSTRUCTION
                                  -COMBUSTOR SECTION (60 IN. LONG X 20 IN. DIA)-

I
in

-------
                                                   SWIRL BURNER ASSEMBLY
I
8
                                                                        SWIRL VANE ADJUSTMENT/INDICATOR
                                                                                                      STATIC PRESSURE
                                          SWIRL VANE
                                         ARRANGMENT
                                                                                STATIONARY VANES
                                                                               MOVEABLE VANES
                                                                                                             COOLING WATER
                                                                                                                         Cxi

-------
                                  TABLE II-B
                   OPERATING CONDITIONS FOR THE IFRF BURNER
     Test Condition
     mair
     P (psia)
     Tinlet
     * *fuel (g/sec)
1, 4+
66.7
14.7
290
3.40
0.8
2. 5+
66.7
14.7
290
4.26
1.0
3, 6+
66.7
14.7
290
5.11
1.2
* „
  Gaseous propane




+ Swirl number (see definition in text) was 0.63 for test  conditions  1-3


  and 1.25 for test conditions 4-6.
                                       II-7

-------
recirculation; however, at higher swirl intensities a recirculating zone  in  the
central portion of the jet is required to support a strong adverse pressure
gradient along the axis.  Since recirculation zones tend to stabilize  the
flame and increase the intensity of reaction, the ability to achieve stable  flame
conditions only above swirl numbers of 0.6 in this research program is  in  agreement
with Bee'r and Chigier's analysis.

II. C. 2  FT12 Burner Can

     A modified FT12 combustor can used in this program was 29.5 cm in  length
(11.5 cm shorter than the original can) and 13.0 cm in diameter.  It was
altered to make all air addition holes symmetric.  This can was welded  at  its
exit to a shroud and was placed in the test section with the swirl burner
removed.  As  shown in Fig. 11-^, a flow straightener was placed in the  burner
housing upstream of the combustor can and appropriate fuel lines and cable for
spark ignition were fed through the housing.  A standard fuel nozzle for the
FT12 (Pratt & Whitney Part No. 525959) was used in this series of tests.

     Three flight conditions, i.e. idle, cruise, and maximum continuous, were
simulated.  A simulation was necessary since the test section could only
sustain a maximum pressure of four atmospheres, while the cruise and maximum
continuous flight conditions required pressure above 6 atmospheres.

     In addition, since gas sampling and accurate definition and examination
of the optical path is simplified by operating at one atmosphere,  all experiments
were performed at one atmosphere.  Simulated flight conditions at this  lower
operating pressure, were calculated by equating Mach numbers.  This is  a common
test procedure and is useful in simulating equivalent fluid flow patterns  and
heat transfer.  In this case, the mass flow rate was reduced appropriately to
maintain the  chamber pressure at one atmosphere and the inlet temperature  was
identical to  the flight conditions.  The simulated flight conditions are listed
in Table II-C.  The optical and probe axis was 78 cm downstream from the exit
of the FT12 burner can.

II. C.  3  Temperature Measurements

     Exhaust   temperatures from the IFRF and FT12 combustors were made using  a
water-cooled, double shielded, aspirated thermocouple probe with a bead made of
Pt/Pt-13%Rh.   The probe was manufactured by Aero Research Instrument Co. (Part
Number T-1006-6 (25) R) according to specifications described by Glawe, et al.
(1956)  but was modified (by water cooling and material substitution) to increase
the temperature range to above 1370 K (2000°F).  .  ohotograph of this probe
is shown in Fig.  II-5.  Radiation and conduction corrections were made  according
to equations   supplied by the manufacturer.  For the measurements made  in this
study (P = 1  atm, Mach no. « 1), these equations can be written

                      Tgas U) = thermocouple U) * * TRC
                                       11-8

-------
                                                   FT12 ASSEMBLY
                                                                        -FLOW STRAIGHTENER
     L7
                 o
                 o
                 o
o
o
o
       L
        MODIFIED FT12

         BURNER CAN
                                                                     /    ii

                                                                TJ1
                                                                                                             FUEL
J IGNITER

   CABLE
                                                                                 AIR
to
I

o
I
00
(JV
I
                                                           ~1/3 ACTUAL SIZE

-------
                             TABLE II-C
             OPERATING CONDITIONS FOR THE FT 12 COMBUSTOR
m
 ar
P (psia)
1 inlet
* *fuel (i/sec)




f/a
Idle
485
14.7
335
5.15
.0106
Cruise
463
14.7
515
6.62
.0143
Max imum
Continuous
454
14.7
524
6.92
.0152
 Jet A
                                 11-10

-------
                                        Pt-Pt/13% Rh ASPIRATED THERMOCOUPLE
                                                                                    (WATER OUT NOT VISIBLE)
u>
I
o
I
                                                                                                                     p

-------
where the radiation and conduction correction, ATRC, is given by

                         ATR(, = 0.55 C1 (K)


                         GI (K) - 0.00917 Tchermocouple (K) - 7.136

and M is the Mach number.  For the FT12, no correction was made since
L TRC was negligible « 10 K).  For the IFRF burner, corrections were
typically on the order of 50 K.  Additional description on the operation of
aspirated probes is given by Land and Barber (1954).
II. D.  Sampling Systems

     Two  instrumentation systems were used for measuring the products of
combustion.  The first is the Scott Exhaust Analyzer used in the Task I of this
program and the second is a chemiluminescence analyzer made by Thermo Electron
Corporation (TECO).  The Scott system was used during the initial tests with
the flat  flame burner to obtain concentrations of the major species.  The Scott
system was thereafter dedicated to the larger scale combustor tests and the
TECO instrument was used for subsequent measurements of NO/NOX over the flat
flame burner.  The Scott system was dedicated to the combustor measurements  for
two reasons.  First of all, it was impractical to move this system between
facilities and secondly it was determined that the Scott package (under the
conditions of the  flat flame tests) did not satisfy the Federal requirements
for total  instrument response time and could not be easily modified to meet
those requirements.

II. D. 1   Scott Exhaust Analyzer

     The  Scott Model 119 Exhaust Analyzer provides for the simultaneous anal-
ysis of CO, C02, NO or N02, Oj and Cotal hydrocarbons (THC).  The
analyzer  is an integrated system, with flow controls for sample, zero and
calibration gases  conveniently located on the control panel.  The incoming gas
simple passes through a refrigeration condenser (
-------
analyzer was  stainless steel, and was operated  at  a  temperature  of
approximately 1000 K.  A Scott Model 150 Paramagnetic Analyzer is used  to
measure the 02 concentration  in  the gas sample.  Concentration ranges available
with this  instrument were  from 0-1% to 0-25% on  several  scales,  with a  nominal
accuracy of + 1% of  full scale.  A Scott Model  116 Total Hydrocarbon Analyzer  is
used to measure the  hydrocarbon  concentration  in the gas sample.  This  analyzer
utilizes an unheated flame  ionization detection  system to  provide for measurement
of hydrocarbons (as methane)  in  concentration  ranges from  0-1 ppm to 0-10%, with
a nominal  accuracy of + 1%  of full scale.   Output  signals  from the  various
analyzers  are displayed on  chart recorders  and  a digital display.

     The sample line was teflon-lined aluminum.  The typical operating  temperature
was 380 K.  When sampling  from the large combustors, water was removed  from
the sample by two traps cooled to 3° C.  The first was located 6 ft. beyond
the probe  exit and the second was housed in the  Scott analyzer.

     II. D. la  Pumpmig^ Requirements

     Two problems specific  to this sampling/probe  system were encountered.
First of all, the orifice diameter of the macroprobes (2mm) was  sufficiently
large that a  separate vacuum  pump (17.5 cfm) was required  to reduce the back
pressure of the probe to the  very low values ( - l/10th of  an atmosphere)
required in this program.   This  vacuum pump was  attached directly to the probe
via a line one inch  (2.54 cm) in internal diameter and three feet (90 cm) long.
The second requirement was  that, at the reduced  pressures, the pumping  capacity
of the sampling system must be sufficient to deliver flow  to the analytical
instrumentation.  To accomplish  this task,  a MB-301 pump and two MB-118 pumps
(metal bellows) were assembled in a series/parallel arrangement.   These pumps
were in addition to  the two MB-118 pumps in the  Scott analyzer and the vacuum
pump associated with the CLA.  As discussed in Section IV. D. 1,  even with this
pumping capacity the deliverable flow was marginal at the  lowest of probe back
pressures.

II. D. 2  TECO Analyzer

     The Thermo Electron Corporation (TECO) Model  10AR Chemiluminescent  NO/NO
                                                                             X
analyzer was used for the reported data for the  flat flame burner.   This
instrument has a stated minimum  detectable  concentration of 50 ppb and a
maximum limit of 10,000 ppm.  Linearity within any of its eight operating
ranges is given as ± 1%.  A TECO Model 300  Molybedenum NOX Converter was used
for the NOo determinations.   Sample was delivered to this analyzer at atmospheric
pressure by a metal bellows pumps (Metal Bellows MB-118).  The sample line was
12 ft of treated teflon line  (Technical Heaters, Inc.).   The back pressure of the
probe was continuously monitored using a Matheson test gauge (0-760 mm,  absolute).
                                      11-13

-------
II.  E.  Mass Flow Measurements

     The purpose for measuring mass flow rates through the sampling probes was
to provide data for comparison with model predictions.  Using the experimental
arrangement depicted in Fig. II-6, flow rates through probes of three orifice
diameters (75, 635, and 2000 microns) were measured at varying external  tempera-
tures and probe back pressures.  To prevent water condensations, heating  tape
was used between the probe and exit of the mass flow meter.  The Hasting  meters
which were used for these mass flow measurements, operate by siphoning a  small
but constant  fraction of the gas flow and passing it over a series of heated
thermocouples.  Cooling of the thermocouples due to the gas flow is measured by
the meter and is primarily a function of the mass flow and the specific heat of
the gas.  According to the manufacturer, mass flow calibrations made for  one
gas can be related to another gas or mixture of gases by the ratio of specific
heats.

     To accommodate the wide range in mass flow due to changes in orifice size
(m >r d orifice) anc* temperature (m >r l/\/Y), three low pressure drop, Hastings
mass flow meters (ALU-100, ALU-5K, ALU-20K) were used.  Full scale on these
Hastings meters were 100, 5000, and 20,000 seem (.00191, 0.0953, and 0.382
g/sec of nitrogen, respectively).  The reported accuracy is ± 1% of full
scale.  In addition, these units each have pressure drops of less than 0.13
torr at an operating pressure of 760 torr (1 atm) and less than 1 torr at
30 torr.  The transducers were installed in the lines according to the manufac-
turer's recommendations.  The assembled plumbing with meter was checked for
leaks by pressurizing to 100 psig.

     To verify their operation two tests were made.  First, a constant mass flow of
nitrogen was  passed through the transducer while the operating pressure of the
transducer was varied by adjusting the valve between the transducer and a
vacuum pump.  At each constant mass flow, the meters produced a constant
reading (typically + 0.5%) independent of the operating pressure.  In agreement
with the specifications, pressure measurements up and downstream from the
transducer indicated a small pressure drop « 3 torr).  Throughout the pressure
range investigated (100-760 torr), the range of mass flows were supplied  using
critical flow orifices upstream of the transducers.  Typical data are presented
in Fig. II-7.

     The second test was a check on the linearity and the calibration of  these
units.   The experimental set-up was identical to the check for constancy  of
indicated reading with varying pressure and constant mass flow.  Experimental
data are shown in Fig. II-8.  Although the linearity appears to be quite  good
(other than a small deviation with the ALU-5K meter), a noticeable discrepancy
with the calibration was found for the ALU-100 (>/> 7%) and ALU-20K (^ 102)
meters.  For  these tests, the calibrations obtained at UTRC were used.  Less
than 12 change was observed for the mass flow measurements shown in Fig.  II-8
when the lines were heated to 60° C.

-------
                         EXPERIMENTAL SET-UP FOR MEASUREMENTS OF MASS FLOW
EXHAUST
I
VACUUM
PUMP

MFTFR .,, I nvM


»
»
^-\ r^-i i- HEATING TAPE

ix^xj "lUllUuI " ""
MASS
(j\ FLOW /

-------
                  INDICATED READING AT CONSTANT MASS FLOW AND VARYING OPERATING PRESSURE
8
                                               O  ALU-100

                                               A  ALU-100

                                               D  ALU-100
                                                     A ALU-20K

                                                     0  ALU-5K
                    1.0
                    0.8
               LU
_j    0.6
_i
=>
U.
LL
O
z
Q    0.4
               cr
                    0.2
                              — O	-O
 — A

-• —

               — D —  -  — D — ---  -- ---- - -  -- D — ^
                                           I
                                       I
                                            I
I
                                                                                          Mw
                                                                                           (Ng

                                                                                      (C3/SEC)  SCCM

                                                                                      0.00182   953
                                                                                                     0.00144    75.4
                                                                                                     0.193   10.110
                                                                                                     0.0410    2150
                                                                                                     0.00064   33.5
                                100       200       300       400       500       600
                                           PRESSURE DOWNSTREAM OF FLOW METER
                                                           (TORR)
                                                                             700
                                                                                                                           O

-------
              METERED MASS FLOW RATES VS. FLOW METER READING
                                      O  ALU — 20K

                                      A  ALU — 5K

                                      D  ALU — 100
                                                                              FIG. n-8
tn

ID
_i
<
O
O
c/3
   6000
   5000
   4000
        — 100
O  3000
g  2000

cc
   1000
        — 60
        — 40
        — 20
                                                I
                 0.2
                           0.4        0.6        0.8


                                FRACTION OF FULL SCALE
                                                         1.0
1.2
            24000



            22000




            20000




            18000




            16000
                                                                                    m
                                                                                    IE
                                                                                    m
                                                                                    CO
                                                                              14000 co

1.4
                                                                              12000
                                                                              10000 O
             8000




             6000




             4000




             2000




             0
                                                                              79-10-85-10
                                         11-17

-------
                  III.  DESIGN OF GAS SAMPLING PROBES
III.A. Losses of Nitric Oxide  in Sampling  System

     Although this report  focuses  on  probe measurements  of  nitric  oxide  and
associated probe phenomena,  it  is  clear  that  a  gas  sample probe  (and  the
corresponding extraction of  a  gas  sample  from a flame  environment)  is  only one
part of a sampling system.   Since  any  portion of  this  multistep  process  (from
gas sampling to species analysis)  could  cause errors,  it is worthwhile to
review this process.  Typically, the  individual steps  include:

     1.  Extraction of the sample  from the flame  environment without
         perturbations external to the probe  due  to  local temperature  changes
         or catalysis (due to  the  presence of a probe).

     2.  Quenching of the  flame gases  inside  the  probe by rapid  temperature
         and, usually, pressure reduction  without the  occurrence of hetero- or
         homogeneous kinetics.

     3.  Removal from the  flame environment and transfer of the  sample to the
         instrumentation without condensation or  reactions  on walls.

     4.  Water removal using,  for  example, an ice trap to minimize condensation
         and/or interference in the detectors yet without condensing or absorbing
         other species of  interest.

     5.  Filtering of the  gas  sample  for  particulates.

     6.  Pressure recovery using a non-interfering pump  to  produce gas samples
         at pressures required  by  the  detector(s).

     7.  Analysis of the sample for the  species of  interest with known or
         calculable corrections for the  presence  of  interfering  species.

     In specific cases, certain items  such as the water  trap, filter, or pump
may not be necessary due to  combustor  conditions, instrumentation, and required
pressures; while in other  systems,  additional facilities, such as  a storage
capability, may be needed.

     In regard to the sampling  and analysis of  nitric  oxide (NO) or total
nitrogen oxides (NO and NOo), nearly  all  of the above  steps have been  sus-
pected and examined as a source of sample  perturbation.  For this  report,
problems associated with sampling  both NO  and N02 are  important  due to the
known interconversion between  these species.  Although many authors have
discussed selected problems  associated with the measurement of nitrogen oxides,
                                      III-l

-------
perhaps  the most comprehensive reviews have been written  by  Cernansky  (1976)
and Tuttle, et al.  (1973).  Complications with measurements  of  nitric  oxides
such  as  those mentioned  in  these reports and with measurements  of  other  gaseous
emissions  led the Federal government  to write regulations (according to  recom-
mendations by the Society of Automotive Engineers (E-31 Committee))  for  the  gas
sampling and measurement of aircraft  emissions (Federal Register;  1973 and
1976).   In spite of  these efforts, many uncertainties  in  the measurements of
nitrogen oxides remain.  Primarily, these uncertainties include:

      1.  Interconversion between nitric oxide and nitrogen dioxide within
         the sampling probe or sampling line.

      2.  Chemical reactions within the probe that reduce  nitrogen  oxides to
         molecular nitrogen or other  nitrogeneous species.

      3.  Sampling line losses of NOY  (i.e., NO and N09).
                                   A                 ^

      4.  Improper calibration or corrections for a chemiluminescent detector
         for the presence of species  other than nitrogen  (the usual diluent in
         calibration gases).

      5.  Low efficiency  for the NO- * NO converter or complete  reduction of
         nitrogen oxides to molecular nitrogen or other nitrogeneous species  in
         the absence of oxygen.

      Experimenters can also observe apparent losses of nitric oxide due to a
variety  of experimental problems, including unconditioned sample lines, small
leaks, and even under unusual operating conditions (e.g., flow  rate through
probe  is less than that required by the analytic instruments) reverse  flow
through  a bypass valve that may dilute the gas sample.  The  above items are
discussed  in detail  in the  following  sections.

III.A.I  N0/N02 Interconversion

      Prior to the early seventies, it was believed that very little nitrogen
dioxide was formed during combustion  processes and that nitric  oxide made up
nearly all of the emissions of nitrogen oxides.   Since that  time, however,  many
experimenters (e.g.  Anon, 1971; Schefer,  et. al., 1973; Merryman and Levy,
1974; Allen, 1975; Kramlich and Malte, 1978; Amin, 1977; Cernansky and
Singh, 1979; Johnson, et. al., 1979,  and Clark and Mellor, 1980) have
probed various combustion systems for nitrogen oxides and have  found large
NO^/NO ratios.   Throughout this decade the source(s) of this measured N07
has been questioned.  Although the flame,  probe,  and the  sampling line have
each been suspected  as its source, it is apparent now that each system must be
analyzed separately.  In a gas turbine combustor, for example,  nitric oxide
formed in the primary zone may be converted to N02 by relatively cold air
entering from the dilution holes (Chen et. al.,  1979).  Alternatively,  as the
                                      II1-2

-------
gases  in  a  probe  are  cooled  to  approximately  1000K,  flame  radicals  are
quenched  and may  be converted to  the hydroperoxyl  radical  (H02)  which  can
oxidize nitric oxide  via  the reaction

                              NO  + H02 -» N02  +  OH                (III-1)

Kinetic analyses  by Johnson, et.  al. (1979) and Kramlich and Malte  (1978)  for
cooled probes  indicate  that  this  reaction is  of prime  importance in the  con-
version of NO  to  N02.   The reaction

                             NO + 0 + M  -»N02 + M                 (III-2)

although  considered, is  relatively unimportant due  to  the short  lifetime  of
atomic oxygen.  In  fact,  the model by Johnson,  et. al.  predicts  that  if  all the
NOy begins  as  N02,  some will be converted to  NO.   In  general,  they  conclude
that for  these flames  (producing  small quantities  of  NOj,,  ^ 10  ppm) , no
relationship exists between  the measured  N0/N02 ratio  and  the  actual ratio  in
the flame.

     Schefer,  et. al.,  (1973) observed the unusual result  that  cooled  probes
(quartz and stainless  steel) indicated virtually no nitric oxide but several
ppm of NO-?  in  an  opposed  jet combustor (premixed,  propane) and with  uncooled
probes the  nitrogen oxides were composed  nearly entirely of NO.   Since the
probes were placed  in  the reaction zone  and sampled only partially  burned
gases, the  authors  argued that  within the uncooled probes  exothermic reac-
tions  continued,  thereby  heating  the surfaces to feed  the  catalytic  conversion
of NO- to NO (similar  to  NOV converters  used  with  CLA).  They  concluded, there-
     /                     A
fore,  that  the very high  N02/(NO  + N02)  ratios  (nearly  1)  obtained  with  the
cooled probes  are realistic measurements.  Based on the recent  studies by
Johnson,  et. al.  (1979) and Kramlich and  Malte  (1978),  it  seems  more reasonable
that Reaction  (III-l)  is  at  least partly  responsible  for conversion  of NO to
N02 in the  cooled probes, especially in  light of the  presence  of oxygen  and
unburned  fragmented hydrocarbons  which are known to produce the  H02  radical
during decomposition.   Any N02  similarly  formed in the  uncooled  probes would
undoubtedly be reconverted to NO  on the  hot surfaces.

     Other experimenters  have considered  the  N02 to NO  conversion (based on
the same  principle as  a catalytic NOX converter) to be  important  for uncooled
stainless-steel probes.   Benson,  Samuelsen, and Peck  (1976) and  Benson and
Samuelsen (1976,  1977), for example, have examined a  similar phenomena in
simulated (heated) probes and in  the presence of carbon monoxide, hydrogen, and
unburned  hydrocarbons.  Their work cannot  be  directly  applied  to probe behavior
since no  (overall) kinetics were  derived  from their work and, more  importantly,
the residence  time  (~ one second) in the  simulated probes  was much  longer than
expected  residence times  in an  uncooled  probe.   This  data  is much more
descriptive of the behavior of  catalytic  converters.
                                       III-3

-------
     In the case of either the NO to N02 oxidation via Reaction (III-l)
or the surface reduction of N02 to NO, neither mechanism can be used for
quantitative predictions at the present time due to the inability to describe
accurately and simultaneously the fluid dynamics, heat transfer, and chemical
kinetics (both hetero- and homogeneous) occurring within a probe.  (Note: The
present study contributes substantially to the understanding of the fluid
mechanics and heat transfer for a certain class of sampling probes.  Section
III.B.2)

     Another possible mechanism for conversion of NO to N02 is the reaction

                             NO + NO + 02 * 2N02               (III-3)

that may occur in the sample transfer lines or in the instrument lines leading
to the reaction or measurement chamber.  Although Cornelius and Wade (1970)
have concluded that this reaction was unimportant in their system, it should be
noted that each sampling system should be examined since this reaction is a
strong function of the nitric oxide concentration and the total pressure.  For
systems with low concentrations of nitric oxide « 250 ppm),  low oxygen concen-
tration, or low sampling line pressures and residence times,  this reaction is
undoubtedly insignificant; however, for systems with large NO concentrations
due to seed NO or fuel nitrogen, with high oxygen concentrations, or with high
sampling line pressures even in only part of the system, this reaction may
convert substantial fractions of the nitric oxide to nitrogen dioxide.   Although
this phenomena was observed in the first phase of this program (Dodge,  et. al.,
1979), complete details of this conversion were not reported.  Since this
reaction is also of interest in this part of the program, further details of
these measurements have been given in the following section.

     Since the rate constant for Reaction III-3 is well known, the contribution
of this reaction can be estimated once initial NO and 02 concentrations are
identified, and pressures and residence times throughout the  sampling system
are measured.

III.A.la  The Bodenstein Reaction

     The reaction

                             NO + NO + 02 + 2N02               (III-3)

under certain conditions may contribute to conversion of NO to N02 in a
sampling system.  It is undoubtedly not a three-body (or termolecular)  reac-
tion but rather represents a sequence which either forms the  dimer,

                                 NO + NO t (NO),                  (III-4)
                                      III-4

-------
                                 (N0)2  +  02  +  2N02

or one  that  forms  the  nitrate

                                  NO  +  02 .>.  N03                     (III-6)

                                   N03  +  NO  +  2N02                  (IH-7)

Although  this has  been discussed by  many authors  (see  review by  Baulch,  et.
al.,  (1970)) the  actual  route  is not resolved.  Nevertheless,  this  reaction  has
been  examined by many  experimenters  and  its rate  constant  is known  better  than
+ 50%, kli;[_3 =  1.2 x  109 exp  (+523/T)  cc2/mole2-sec  (Baulch,  et.  al.,  1970;
Hillard and Wheeler,  1977)  when  the  reaction  rate  is defined as  in  Equation
III-8.

      The  rate of  loss  of NO with respect to time  due to  this reaction  can  be
written

                           d[NO]                 9
                            dt   = -2km_3  [NO]-*  [02]              (III-8)

where brackets,  [  ], represent concentration  of the molecule in moles/cc.
According to this  equation, the  rate of  the Bodenstein reaction  is  dependent on
the square of nitric oxide  concentration and  the  first power of  the  oxygen
concentration.   Since  number  densities are  directly  proportional  to  pressure,
this  reaction rate is  also  indirectly  dependent on  the total  pressure  to the
third power.

      This  reaction then  becomes  quite  important in  sampling  systems  where
either high concentrations  of  nitric oxide  or oxygen exist,  high  pressures
exist, or  long  residence times occur.  Alternatively,  their  opposites  will tend
to disfavor Reaction III-3.   Specifically,  this reaction may  be of  importance
when  NO or fuel nitrogen is added to the flame  and when high  sample  line
pressures  exist, even  for  short  line lengths.

      During the course of  the  Task I investigation  (Dodge, et. al.,  1979)  high
concentrations  of  NO were  added  to a H2/02/Ar flat  flame so  that  infrared
optical measurements could  be made.  Simultaneous probe measurements were  also
made during these  tests.   For  the particular  sampling  conditions  used  in these
experiments,  Reaction  III-3 was  found  to contribute  significantly to the
conversion of NO to N02.  The  results  can best  be understood  by examining  the
complete  sampling  system.   In  Table  III-A,  a  review of different  components  of
the sample system  from the  probe  tip to  the reaction chamber  in the  CLA  is
given.  Estimates  of lengths of  line,  local pressure,  residence times, and
temperatures  are provided.  This  sampling system is only part  of  the complete
                                       III-5

-------
                                                          TABLE  III-A
                                           FLOW  CONDITIONS  IN  NON-IDEAL  SAMPLING  SYSTEM
                                                                                                              Reaction Chamber
               Probe and Sample
                Transfer Line
                                      Refrigerator
                                          (2)
                           Metal Bel-
                           lows Pumps
                       Flowmeter and Con
                         nection to CLA
                                                     H20  Re-
                                                     moval
                                                                                Bypass
                                                                                Valve
Distance
cm (feet)
                  550 (18)
1070 (35)
460 (15)
470 (15.5)
Pressure
torr (atm)
                  350 (0.46)
 350 (0.46)
785 (1.03)
760 (1.0)
Eat imated
Residence
Time (sec)
                   1.6
   4.2
  3.7
  3.7
Temperature
   CO
                  110
                                                                    20
                                                       20
Conditions for
                        1600 K

-------
SCOTT  Instrument  system  in  which  the  gas  can be  transferred to any of five
analytical  instruments.   In a more compact  system containing only a
CLA, it  is  expected  that  line lengths and therefore residence times will be
noticeably  shorter.   For  the estimates provided  in Table III-A,  the individual
residence times were based  on total residence times and estimated mass flow,
temperature,  cross-sectional area, etc.  in  each  section.  The total residence
time was measured  as the  time betweeen the  moment when NO is first visually
observed after a  toggle valve is  opened  to  add seed NO to the flame (i.e., when
the  flame turns a  greenish-gray color) and  the moment  when the indicated nitric
oxide  concentration  begins  to rise rapidly.  The pressures at various locations
in the  sampling system were obtained  by  placing  several pressure gauges  along
the  sample  lines.  The sampling system was  operated to meet the  Federal  Require-
ments  for a sampling system with  the  following exceptions.  First of all,
the  sample  line was  held  at 110 °C rather than the required 150  °C since,  in
this case,  the sampled gas  consisted  of  1^2/^2 combustion products and
condensation  of hydrocarbon fragments was not of concern.   Secondly,  the total
response time of  the analytical system (from probe tip to  15% response at  the
detector as defined  by the  Federal Register) was approximately 18 seconds  vs.
the  required  time  of less than nine seconds.  For sampling over  the flat flame
burner with  the SCOTT instrument  package, this time could  not  be  greatly
reduced at  the flame temperatures  of  these  measurements (
-------
                               = 2kTTT »[09] t- +  1                (III-9)
                                      '
where lNO]Q is the nitric oxide concentration before passing through
the ith section of the sampling line and [NO]^ is the concentration at the
end of this section.   By summing Equation III-9 for each section (accounting
for pressure variations appropriately) and assuming that the measured NOX
estimates the original nitric oxide concentration at the probe tip, the final
value of NO just prior to the CLA may be calculated.  These values were esti-
mated for several experimental seed levels of NO at various flame conditions.
Calculated and experimental values of NO and N02 are reported in Table III-B.
Good relative agreement is found between the calculated and measured values of
N02 (see ratios) although the calculated value is approximately a factor two
higher than the measured value.  This factor of two difference is not due to
an error in the rate constant but is probably due to inaccurate estimates of
the flow parameters in the sampling system and/or loss of N0~ in the system.
In any case, the similarities in the trends and magnitudes of conversion
between the calculated and measured values of N02, provides good evidence that
Reaction III-3 is responsible for the conversion of NO to N02 in this sampling
system.

III. A. 2  N0x Reduction in Sampling Probe

     For uncooled stainless-steel probes sampling fuel-rich flames there
exists much evidence that nitrogen oxides can be reduced (probably to molecular
nitrogen) within the probe.  This phenomena is not unexpected since it is the
same as that observed when attempting to use a stainless-steel catalytic
converter to convert N02 to NO in the presence of fuel-rich gases.  Reduction
by stainless-steel probes has been observed by Halstead, et. al. (1972) who
compared an uncooled quartz-lined probe with an uncooled stainless-steel lined
probe and by others (England, et. al., 1973; Cernansky and Singh, 1979; and
this work).  This problem is easily eliminated or at least drastically reduced
by using cooled probes for sampling fuel-rich flames.

     Under stoichiotnetric or fuel-lean conditions very little evidence exists
that indicates probes significantly alter total concentrations of nitrogen
oxides.  England, et. al (1973) have observed some dependence of NO concentra-
tion with changes in probe type; total NO , however, is not reported in this
report and some of the differences may be due to interconversion between NO and
NOo.   In addition, the only major differences are noted between the cooled
and uncooled probes.   Few, McGregor, and coworkers (1972, 1975, 1976, and 1977)
have compared UV optical measurements of NO to probe measurements and have
concluded that the probe measurements are up to a factor of six lower than the
optical measurements.  These conclusions are, however, subject to question due
                                       III-8

-------
                                  TABLE III-B
               COMPARISON OF CALCULATED AND MEASURED NO AND NO,
                                Measured
Calculated
                                                             1
Calc/Meas,
Calculated
°2
%
5.0
5.0
5.0
7.9
7.9
Measured
NOX
(ppm)
4891
7266
7114
3312
4533
NO
( ppm )
4727
6833
6715
3189
4272
N022
(ppm)
164
433
399
123
261
NO
(ppm)
4550
6539
6416
3080
4110
N02
(ppm)
341
727
698
232
423
N02
2.08
1.68
1.75
1.89
1.62
1
 Estimated concentration at CLA assuming the initial  NO concentration at  the

 probe tip is equal to the measured NO
 Measured NO  minus measured NO
            X
                                      III-9

-------
to errors in the calibration procedures and theoretical model (Dodge, et. al.,
1979).  Cernansky and Singh (1978) have observed some differences in total
NOX for a variety of probes sampling fuel-lean, stoichiometric and fuel-rich
flat  flames; however the differences are typically small « 15%).

      Recently, Clark and Mellor (1980) have compared NO and NOX measurements
using several different probes in a model gas turbine combustor.  They report
some  rather large differences between measurements of NO using blunt and
tapered tip probes (as much as a factor of three); however, measurements of
total N0x indicate relatively small scatter (~20%) which is typical of the
day to day variations in their combustor and/or analysis system.  It should be
noted that this agreement is achieved despite high measurements of hydrocarbons
(>4%) that indicate the probe is sampling within the reactive flame zone.

      Optical and probe measurements have been compared at other laboratories
and in general good agreement is found.  Meinel and Krauss (1978) have made in
situ measurements of NO in both H2/air and C^Hg/air laminar premixed
flames using a UV resonant lamp.  In lean flames,  agreement between the probe
and optics is excellent, i.e., within several percent.  For rich flames, the
optics produce values which are not greater than the probe values but are
approximately 20 to 25 percent lower.  Falcone, et al. (1979) have made optical
measurement  using an infrared, tunable diode laser.  For a lean flat flame,
0 = 0.67, the optical measurements are about 20 percent higher than the probe
values; however, they conclude that the primary uncertainties are associated
with  the  laser system since the probe measurements agree well with the seed
values of NO.

      Bilger and Beck (1975) have made probe measurements on a turbulent diffu-
sion, hydrogen/air flame and compared these measurements to those of previous
work  on the same or similar system.  The measurements of the major species,
i.e., Hj, H^O, O^, in general agreed quite well with only a small axial
shift observable.  The NO measurements, however, differed quite noticeably with
the more recent results (using a probe with a slender nose) suggesting peak
NO concentrations as much as 30 to 35 percent higher than earlier measurements
with  a probe having a blunt nose profile.  In addition, they compare data where
substantially different NO profiles between the small and large probes are
observed with the small probe producing the highest NO.  A substantial shift was
also  observed in the major species.  For the large probe, the NO profiles were
dependent on the flow rate, but no difference in the major species were observed,
In these tests, it is not too surprising that different probes produce different
results since the length of the turbulent flame is on the same order as the
probe diameter.  For example, the flame length is  approximately 5 to 6 mm and
the blunt-nosed probe has a diameter of 6 mm very  near to its orifice.  This
large probe  certainly must be considered a poor design for probing the reac-
tion  zone of such a flame.  The slender nosed probe is a better design with an
initial diameter greater than 1.2 mm but even this probe quickly tapers back to
U mm  diameter.  The differences in the profiles (both NO and major species)  may
be due to the presence of a large heat sink, i.e., the probe, or possible
                                      111-10

-------
stagnation zones in  front of these probes.  The difference between  the NO
profiles obtained using the large probe at various  flow rates may be  due to  the
phenomena stated above or to NO/NOj  interconversion within the  probe.  NOj
was measured only for the smaller probe and in  that case was  found  to be
negligible.  Bilger  and Beck also conclude that the slender nosed probe provides
a more realistic measurement of nitric oxide.

     Bryson and Few  (1978) at Arnold Research Organization (ARO) have observed
relatively large differences (^ 50%) in total nitrogen oxides between a tubular
probe and either a  'quick-quench' or dilution probe (the latter  two agree).
Since this discrepancy is rather large, it is worthwhile to investigate the  ARO
study in detail to  see if this report  identifies  areas that require further
research.

     Quite clearly,  there are several  significant conditions  of  the Bryson and
Few study that are  different from other reports.  First of all,  in  their study
supersonic exhaust  from an AVCO-Lycoming engine is  sampled.   In  all other
papers that report measurements using  different probes, subsonic flows are
examined.  Secondly, three distinct, water-cooled probes were compared:  a
tubular inlet probe, a 'quick-quench1  probe, and  a  dilution probe.  In addition,
the stainless-steel  tubular probe was  constructed to accept inserts of copper
or fused silica.  Alternatively, other studies  primarily examined effects due
only to changes in  surface material.   (Some papers, e.g., England, et al. (1973),
do test some modifications to probe  design but  in general these  changes are
minor relative to the design variations found in  the ARO study).  In  regard  to
the design variations, it should be  noted that  the  tubular probe which produced
the relatively high  NOy measurements also had a very large opening  (0.77 cm
i.d. vs 0.12 and 0.127 for the other two probes)  and was operated with a sample
line pressure much  larger than the latter two probes.  The ratios of  sample
line to combustor pressure were typically 0.9 for the tubular probe and less
than 0.5 for the other probes.  Undoubtedly the pumps used for  the  'quick
quench1 and dilution probe were insufficient to choke the tubular probe.  Under
choked conditions (not well defined  for a constant  area tube where  the flow
should friction choke at the exit rather than at  the entrance),  the probe would draw
approximately 40 times the flow of the smaller  probes.  Based on the  above
analysis, it seems  likely that the different NO^ measurements between a
tubular probe and either a "quick-quench1 or dilution probe may  be  associated
with differences in  sampling pressure  and flow  rate through the  probe and/or
the existence of a  stagnation zone in  front of  the  probes.

     A stagnation zone, for example, in front of  these fairly blunt probes in
a supersonic stream  could perturb the  gas samples.  For air at  a Mach number
of 1.15, for example, a stagnation of  the flow  results in a 25%  rise  in temper-
ature.  For the tubular inlet probe, the problem may be the most severe.  First
of all, the gas decelerates to a very  low Mach  number at the  entrance (estima-
ted to be less than  0.1) and consequently stagnation temperatures must be
                                      111-11

-------
approached even in the absence of an external stagnation zone.  Secondly,  the
rate of cooling is undoubtedly much slower in this probe than  in  the  other two
since the sample tube is quite large (producing a low surface  to  volume  ratio
for heat transfer) and since the pressure and mass flow through this  tube  are
significantly larger.  The problem should be less important when  sampling  ex-
haust gas at lower temperatures because the magnitude of the temperature rise
is less and since kinetics are slower at lower temperatures.   In  Bryson  and
Few's study, essentially no difference between the probes was  found for  the NO
measurements at the lowest stoichiometries which produce exhaust  temperatures
approximately 600 K less than the other tests.

     Sample line leaks from either the atmosphere or from a purge system may
also affect relative readings when a probe and/or pumping system  create  large
differences in sample line pressures and flow rates.  Consider a  case with a
small leak across a purge valve.  If the sample line pressure  is  only slightly
less than the purge pressure and the mass flow through the sample line is  large,
the leak would be unchoked and, if small, would dilute the sample minimally.
However, when the sample line pressure reduces to below 50% of the purge pres-
sure and simultaneously the sample flow rate decreases (such as by reducing
the probe orifice diameter) then the importance of the leak may increase sub-
stantially.  Based on the data presented by Bryson and Few, the above phenomena
could explain some of the differences between the tubular inlet and the  other
two probes.  This effect, however, is not believed to be the cause of the  dis-
crepancies since at low power levels NO measurements did not vary with probe
type.  It is more likely that the increase in temperature (and pressure) due to
shock recovery and subsonic diffusion created the sampling problems encountered
by Bryson and Few.

     Another aspect of the study by Bryson and Few is that, for the dilution
probes,  wide scatter was observed.  The authors comment that part of  the scatter
may be due to estimation of the dilution ratio which was calculated via  two
techniques.  A factor not considered in their study is that a  change  in  the
diluent  or carrier will change the response of a chemiluminescent detector.
The response is not only a function of the quenching efficiency of a  third  body
as pointed out by Matthews, et al. (1977) but also dependent on the viscosity
of the sampled gas for standard commercial units (Folsom and Courtney, 1979  and
Dodge,  et al. , 1979a).   Although this phenomena is discussed in greater  detail
in Section III.A.4, it  can be estimated based on data from Folsom and Courtney
that the NO., measurements made when diluting with argon should be increased
by approximately 10 percent depending on the diluent ratio.  No data  were  found
for the  case of a helium carrier, but since the viscosity of helium and  nitrogen
are similar and if helium is less efficient than nitrogen for  the reaction

                              N02* + M * N02 + M                  (111-10)

which competes with

                                N02* + N02 + hv                   (III-ll)
                                      111-12

-------
then one would expect that the data obtained with  a diluent  of helium  should be
reduced by several percent.  We estimate  that  these corrections  should reduce
the uncertainty for data obtained using the dilution  probe.

     Additional research that may indicate that  probes  perturb measurements of
total nitrogen oxides has been performed  by England,  et  al.  (1973); however, in
this work, only nitric oxide data are  presented.   Consequently,  definitive
statements on the  loss of total NOX cannot be  made.   For  uncooled,  stainless-
steel probes in rich  flames, the NO measurements are  significantly  less  than
for cooled probes.  As described previously this behavior  is not  at all  un-
expected for 0 >  1.0  (rich flames).   In spite  of the  fact  that the  paper
indicates a fall-off  in NO for these  uncooled  probes  even  for 0.6 < 0  <  1.0,
this may be an artifact of the curve-fitting technique  (no  actual data points
are published).   In any case, no detailed mechanism has  been presented to
explain these data.   For two cooled probes and even an  uncooled  quartz tube, NO
profiles are nearly identical for the  lean flames  and are  similar for  fuel-rich
flames.  Slight discrepancies on the  fuel-rich side may  be  due to NO/N02
interconversion rather than  loss of total NOy.

     In summary,  there is only one known  report  that  suggests that  changing probes
may change the total  concentration of  nitrogen oxides by  a  significant amount
(other than uncooled  probes  in rich flames).   This study by Bryson  and Few
shows that a tubular  inlet probe produces higher values  (by 50%) than  two other
probes.  Although  insufficient data are available, it is  possible that pheno-
mena associated with  the higher operating pressure, an  external  stagnation
zone, or different flow rates may contribute to  the observed results.  Moreover,
it should also be  noted that apparent  differences  in  NOy may be  observed
between different  probes if  NO/N02 interconversion varies between the  probes
and N©2 exiting from  one of  the probes is lost in  water  traps or an ineffic-
ient converter.

III.A.3  Losses in Sampling  Line

     As Tuttle, et al. (1973) have discussed,  NOo  can be  lost in sampling
lines for any of  various reasons.  These  include reduction  of N0£ on parti-
culate filters, loss  in water traps,  or loss or  reduction of N02 in uncon-
ditioned stainless-steel sampling lines.  Dimitriades (1967) claimed to  have
observed NOo losses in both  water traps and Drierite  columns, but these
losses are not quantified and it consequently  is difficult  to estimate the
impact on a measurement where water removal prior  to  analysis is necessary.
Dimitriades does,  however^ comment that this loss  is  specific to N©2 and that
he did not observe similar losses of  NO.  The  same comment  cannot be made for
the other items listed above.  For example, it has been  observed in our  labora-
tory that unconditioned stainless-steel lines  will also  produce  an  apparent
loss of nitric oxide.  This  loss mechanism can usually be eliminated by  flowing
an NO calibration gas through new lines for 15 minutes  to  1/2 hour.  Once
conditioned, no evidence was found that indicated  the necessity  to  repeat this
process.   In regards  to the  loss of N02 on filter  paper  containing  carbon or
soot particles, Tuttle, et al. observed losses of  N02 using an NDUV detector
                                      111-13

-------
(for N02) but did not look for the possible conversion to NO or for similar
losses of NO due to the presence of soot.   Gas-phase, NO-hydrocarbon reactions
are typically as well known and as fast as N02-hydrocarbon reactions; by
analogy, one might expect similar reactions between NO and deposits of soot.

     Each of the above problems represents a real concern and should be con-
sidered in the construction of any sampling or analysis system.  Presumably,
the line conditioning problem can be easily rectified by using teflon coated
tubing and fittings throughout or using clean stainless-steel tubing which has
been conditioned for several (> 15) minutes using an NO or N02 calibration
gas.  Nitric acid should not be used as a cleaning agent since it leaves
nitrates and/or nitrites as residues.

     Water condensation is a more difficult problem to solve.  Bryson and Few
(1978) ran without a water trap and then between measurements flushed out the
lines with dry nitrogen in the analyzer to remove any condensed water.  (This
procedure complicates data reduction since, as discussed Section III.A.4,
water can have a strong effect on the response of the CLA and was not accounted
for by Bryson and Few.)  Although this technique eliminates the retention of
large quantitites of water within the analyzer some condensed water along the
inner walls must remain and presumably could contribute to NOo absorption.
The best methods would include the modification of the analysis equipment so
that (1) the entire system (both sampling and analyzer) are heated substantially
above the dew point (based on the maximum pressure in the sampling system and
the initial water concentration) or (2) sample line and analysis instruments
all operate at reduced pressure (low enough to prevent condensation).  For
example, for a liquid-fueled gas turbine engine operating at an overall stoi-
chiometry less than 0.33, it can be estimated that all lines should be heated
above 35°C if the maximum sample pressure is only 800 torr (15.5 psia).
Although these modifications represent changes over the present specifications
in the Federal Register (1973, 1976), it is not clear whether such steps are
necessary since the loss of N02 in a trap is not understood quantitatively
and may be dependent on trap geometry and capacity; among other things.

     The loss of NO,, due to the presence of soot on particulate filters (or
even coated on sample lines) is a more difficult problem to address.  Presumably,
no easy solution can be found for a combustion system that produces large
quantities of soot.  Fortunately, Federal requirements for smoke emissions help
to reduce potential complications due to soot in sampling lines.  As in the
above case for water absorption, losses cannot be estimated quantitatively.

III.A.4  Response of Chemiluminescence Analyzer

     As mentioned previously in Section III.A.2, the response of a commercial
chemiluminescent analyzer (CLA) is dependent not only on the concentration of
nitric oxide but also on the fluid mechanical properties and quenching effi-
ciency of the carrier gas.  Response changes due to changes in quenching
efficiency have been reviewed by Matthews, et al. (1977).  From the well-known
chemiluminescent reaction sequence
                                      111-14

-------
                               NO  + 03 -» N02*  +  02                      (111-12)

                                         -»N02 + 02                     (111-13)
                                  *
                               N02   +  M -> N02  +  M                      (111-10)
                                  *
                               N02   -*N02  +  hv                         (III-ll)


 it  is  clear  that Reactions  111-10  and  11  are  in  competition.   The  presence  of
 species  such  as water  and  carbon dioxide  which are  more  efficient  than nitrogen
 at  quenching  the excited N02 molecule  (N02  )  will produce  less of  a
 response  than when  nitrogen is the  carrier  for equivalent  concentrations  of
 nitric oxide.  The  presence  of species  that are  less  efficient at  quenching
 (e.g., argon) correspondingly  will  produce  a  greater  response.

     These comments  are true only  in the  case of a  CLA for which the  pressure
 in  the reaction chamber is  maintained  constant (such  as  the  original  design by
 Fontijn).  Unfortunately,  the  technique used  to maintain constant  pressure  in
 the reaction  chamber of low pressure commercial units  is via  flow  restriction
 using  capillary tubes.  Since  the  flow  rate through these  tubes  is dependent
 upon on  the  fluid mechanical properties of  the carrier,  the  pressure  and, of
 course,  the  concentrations  of  NO and the  quenching  species (M)  in  the  reaction
 chamber  are  dependent  on these same properties.  This phenomena was observed
 several years ago by one of these  authors (M. F. Zabielski) when comparing  mass
 spect rometr ic and chemiluminescent  data when  argon  was used  as  the carrier.  In
 this case, the chemiluminescent detector  was  found  to have a  reduced  response
 relative  to measurements with  nitrogen  carrier; this  direction  is  opposite  to
 that predicted according to an analysis of  quenching  phenomena.  Subsequently,
 Dodge, et al. (1979a)  modelled this problem by assuming  frictional choking  in
 the capillary tubes.   Stimulated by the original UTRC work, Folsom and  Courtney
 (1979) performed a  detailed  empirical  study on the  effect of  carrier  gas  on  the
 responses of  commercial Beckman and Thermo  Electron instruments.  Although
 Folsom and Courtney  qualitatively  explained their data in terms of relative
 viscosity and quenching data,  the model developed at UTRC (Dodge, et al .  1979a)
 appears  to be in good  quantitative  agreement with the data obtained both  at UTRC
 and by Folsom and Courtney  on  the Thermo Electron instrument.  (The UTRC model
would have to be modified  for  the  atmospheric pressure instruments (e.g.  Beckman
or McMillan) due to a  different design.

     In general, corrections due to differences in a carrier  of pure nitrogen
and that of combustion gases are not large  (typically < 10%).  Exceptions to
this rule include flames using an unusual carrier (argon, for  example)  or no
carrier at all; samping systems where the water is  not removed (especially  for
stoichiome tries near $ = 1.0); or samping systems were the sampled gases  are
diluted with  a species other than nitrogen.   In these cases,  the combination of
viscous and quenching  effects  ought to be considered, since the response  change
of a CLA can amount to as much as 15% or -possibly more.
                                      111-15

-------
III.A.5  NOX Converter

     The purpose of an NOX converter is to reduce nitrogen dioxide to nitric
oxide via the overall reaction

                                2N02 — 2ND + 02                     (111-14)

in the presence of a heated metallic surface (usually stainless-steel).  The
NO thus formed, along with the initial NO,is detected in a chemiluminescent
analyzer.   The N02 concentration is obtained by subtracting from this response
the instrument response when the converter is bypassed.

     Problems associated with the operation of the converter have been discussed
in detail by Tuttle, et al. (1973).  The problems can be placed into the three
classifications:  anomalies, inefficiencies, and NOy reduction.  Tuttle, et
al., for example, report an anomaly observed by A. Nelson of Pratt and Whitney
Aircraft.   In this case, a span gas of 91 ppm NO in nitrogen but void of N02
(verified by the Saltzmann technique) indicated a concentration of 95 ppm when
passed through the detector.  As opposed to a decrease or apparent loss of NO,
this increase appears very unusual since it suggests a generation of nitric
oxide.  Although Tuttle reports this as an anomoly, it is likely that the
calibration of the instrument shifted (if the operator did not adjust for
changing flow rates due to the pressure drop across the converter) or perhaps
nitric acid was used as a cleaner which may outgas nitrogeneous species.  It is
more typical that a slight but noticeable reduction in indicated NO (several
percent) will be observed when a span gas of NO is directed through the converter,
Presumably this loss is due to a reduction of nitric oxide (to nitrogen) in the
converter.

     Practically, the best efficiency for conversion of N02 to NO is about 97
to 98 percent, and efficiencies of 90 to 97 percent are typically achieved for
fuel-lean gases.  These efficiencies, however, are dependent on NO,, concen-
trations and type, condition, and temperature of the converter.  Most converters,
for example,  will perform, at least momentarily, when fuel-rich gases with
NOX are passed through, however, stainless-steel converters at high tempera-
tures will last only a matter of seconds before total NOX is destroyed.
Molybdenum converters are more useful since they efficiently convert N0« to
NO without loss of total NO,, for up to a minute or so (depending on flame
stoichiometry and converter temperature).  The increased activity of molybdenum
for fuel-rich gases is due to its lower operating temperature ( - 450°C for
molybdenum vs.  - 700°C for stainless-steel).  The activities of the converters
are easily recovered by flowing an oxygen rich mixture through the heated tubes
for a few minutes.  In fact, alternately flowing air through the converter when
sampling fuel-rich flames is a common procedure.  This technique should be used
with caution, however, since resultant surface oxidation deteriorates the
converter and may reduce its conversion efficiency.
                                     111-16

-------
III.A.6  Summa ry

      In the above discussion,  it  is  shown  that many  phenomena may  contribute  to
erroneous measurements of nitrogen oxides.   These  include  interconversion
between NO and NCU, loss of  total NOy, and misinterpretation of data
(assuming use of CLA).  Interconversion  and  loss can occur  in the  probe, sample
line  (water trap and  filter  included) or converter.   Perhaps the biggest errors
may be caused by the  use of  uncooled probes  sampling fuel-rich gas.   Under  the
same  flame conditions, NC^ measurements  (using a converter/CLA) can only be
made  through careful  use of  the converter.   In any case, efficiency checks  on
the converter operation should always be made.  In the  case of lean or  stoichio-
metric flames, only one report (Bryson and Few) was  found  that indicated
substantial discrepancies (~50%)  in NOX when measured using different water-cooled
probes.  These discrepancies may  be  due  (at  least  in part)  to differences in
sample line pressure  and/or  flow  rate.   The  results  of  another report which
indicates discrepancies (Bilger and  Beck)  are unreliable since at  least one of
the probes was poorly matched  to  the combustion system.  The only  other signifi-
cant  probe effect is  the conversion  from NO  to N02 due  to  oxidation by  the
H02 radical.  This reaction, however, has  been found to be  of significance
only  when relatively  low concentrations  of NO are  present  (-50 ppm).   It would
be expected that at larger concentrations  when the NO/HC^  ratio is large both
in the flame and probe, oxidation of NO  via  the HC^  radical is relatively
unimportant.
III.B  Quenching  in Gas Sampling Probes

     Gas samples  extracted  from a  flame environment are quenched by a rapid
reduction in either pressure or temperature, and typically both.  For a bi-
molecular reaction, a reduction in pressure by a factor of ten will reduce reaction
rates by a factor of 100.   Decreasing  the gas temperature from 2000 to 1200 K for
a reaction with an activation energy of 40 kcal/mole  (typical for overall
hydrocarbon oxidation) will decrease the reaction rate by a factor of 750.  The
quenching of flame gases is required not only to stop ongoing reactions (non-
equilibrium conditions) but also to prevent a shift from equilibrium (or
quasi-equilibrium) conditions such as may be present  in a post flame zone.
Ideally, a probe  should quench or  "freeze" the flame  gases exactly at the
concentrations present where the probe tip is located.

     Techniques used to quench flame gases include quenching by dilution,
convection,  or expansion and each  of these have been  discussed in some detail
by Tine (1961).   Briefly, quenching by dilution is accomplished by adding a
low temperature diluent to  the flame gases which acts to absorb heat from the
extracted sample.  Quenching by convection is performed by heat transfer to
cooled walls within the probe and  quenching by expansion is accomplished by
                                      111-17

-------
accelerating the gases supersonically through a nozzle to drop  the  static
temperature and pressure.  Although expansion cooling is undoubtedly  the
fastest (-108 K/sec vs.  - 106 K/sec for cooling by convection),  the
associated phenomena are not well understood and two major drawbacks  accompany
this technique.  First of  all, convection cooling to cooled walls must  also
take place while the gas is flowing supersonically; otherwise,  upon return to
subsonic flow conditions,  the static gas temperature will return  near to flame
temperatures.  Secondly, substantial pressure losses, both friction and normal
shock losses, are suffered.  Details of these phenomena will be discussed  later
in this chapter.

     In spite of the many  years that probes have been in use, there exist many
uncertainties in regard to their quenching behavior.  Beal and  Grey (1953) have
argued that any of the above three techniques provide sufficiently high quench-
ing rates to freeze the concentrations of stable species at or  near concentra-
tions present in the flame zone.  Alternatively, Halpern and Ruegg  (1958) have
found that changes in quenching rates due to changes in probe design, internal
diameter; and sample flow  rate may vary measured ratios of CO/CC>2 and ^7
HoO.  These latter conclusions may be subject to error since the  authors
apparently used probes that were comparable to the size of the  burner.  The
presence of a large, cooled probe will significantly perturb the  flame environ-
ment.  Other authors have  reported discrepancies in CO and NO measurements for
various flame conditions using different probes (Bryson and Few,  1978; Bilger
and Beck, 1975; and England, et al., 1973) but it is not clear whether these
discrepancies are due to differences in quenching rates rather  than to  flame
perturbations, stagnation  of the flow, or sampling line/analysis  phenomena.  It
should be noted that species existing in low concentrations (several hundred
ppm or less) are especially susceptible to any of the above effects.  Examples
include NO, N02 and in lean flames, CO.  Small absolute changes (-25-50 ppm)
can reflect a large relative change for these molecules.  In fact, increased
rates of quenching may not necessarily solve problems associated with measuring
these species since radical termination on the walls or in the  gas phase may
perturb concentrations of  these species.  Indeed, a model developed by Kramlich
and Malte (1978) makes the unusual prediction that higher quench  rates  increase
the conversion (in the probe) from NO to N02.

     Not only are there uncertainties in the effect of quenchir?  rates, but also
actual knowledge of quenching rates is unknown.  Typically, approximate fluid
mechanic models are used to estimate cooling rates and pressure reductions;
however, virtually no experimental evidence exists to verify these calculations.
For example,  the operation of quartz microprobes has been misunderstood for
years.   Fristrom and Westenberg (1965) argued by analogy that since a large
diverging nozzle sustains a supersonic flow, a smf '1 probe nozzle should
perform similarly.  Although this extrapolation is Somewhat suspect, many
experimenters have operated on this premise with no experimental  verification
and very little,  if any, theoretical analysis (e.g., Friedman and Cyphers,
1955; Lyon,  et al., 1975; Kramlich and Malte, 1978; Lengelle and  Verdier,
1973).   More recently,  questions regarding these conclusions have been raised
                                      111-18

-------
in the  literature  (Bilge^  1975, Amin,  1977;  Seery,  et  al.,  1977;  Cernansky  and
Singh,  1979) with  fewer  and  fewer  experimenters  tacitly  assuming  the  existence
of quenching by expansion.   Even so,  arguments used  to  question supersonic
expansion  are  primarily  phenomenological with only a few studies  attempting  to
examine analytical details  of  the  fluid mechanics  (Seery,  et  al.,  1977;  Cohen
and Guile,  1970; Amin,  1977) and with virtually  no experimental investigation
of the  problem.

     The  lack  of understanding of  this  problem can be explained by the  fact
that the  fluid mechanic  and  thermal  status  of the  sample gas  in a  probe  is
altered by  a number  of  factors.  These  include:  heat transfer  from the  uncooled
probe tip  to the sample  gas  and  from the sample  gas  to  the  probe  coolant; skin
friction;  flow area  changes; pressure losses  associated  with  shock systems,
sudden  area expansions,  and  turns  within the  sample  passage;  and,  chemical
reaction.   For probes designed to  achieve an  aerodynamic quench (see  III.B.2.a),
sudden-expansion losses,  turn  losses, and chemical reaction  are neglected and
shock losses are avoided  until the aerodynamic quenching region of the  flow  is
completed.  To understand the  influence of  the competing mechanisms,  it  is
convenient  to  assume  that the  sample flow is  both  steady and  one-dimensional
and that manufacturing  techniques  are sophisticated  enough  so that  the  internal
geometry of the probe (especially  in the region  of the  probe  tip)  is  reasonably
close to  the geometry analyzed.  (The wall  temperature  distribution within the
probe tip  is difficult  to calculate  and, for  small diameter  probes, the  internal
shape of  the tip is  difficult  to control during  manufacture  and also  difficult
to examine  once constructed.)

     Even  though a model  may be developed to  describe heat  transfer,  skin
friction,  shock losses,  and  the effect  of area changes,  each  of these phenomena
produces different relative  effects  as  the  probe operating  conditions are
varied.  Consequently,  it is necessary  to examine  each  experimental condition
individually and a probe  designed  from  such a study  is  usually  a compromise.

     The above discussion indicates  the complexities of  the  operation of a
probe in terms of  kinetics,  fluid  mechanics,  and heat transfer.  Due  to  the
discrepancies  between optical  and  probe measurements observed by McGregor, Few
and coworkers, and the  above problems,  the  present program  focused  on understanding
probe and  sample line behavior and its  impact on sample  analysis.   Although
substantial effort was  placed  on examining  possible  homogeneous mechanisms for
NO reduction, no mechanism was found that predicts loss  of nitrogen oxides
within water-cooled  probes.  Interconversion  between NO  and N02 may affect NO
measurements and was observed  in this study under  certain  conditions.  This
phenomena,  however,  was  of  little  importance  since,  in  the  comparisons of
optical and probe measurements, NO.,  was approximately equal  to  NO.  Moreover,
recent studies on  this  problem (Johnson, et al., 1979;  Kramlich &  Malte, 1978)
suggest that total NO,,  is  conserved.  Knowledge  of the N02/NO ratios  may be
important  in understanding  the  total  emissions problem  (N02  can be  visible in
sufficient  concentrations  and  NOo  acts  as an  initiator  in  the photochemical
smog cycle).
                                      111-19

-------
     Fluid mechanics and heat transfer mechanisms were also examined  in  this
study.  Particular effort was focused on the phenomena of aerodynamic  cooling
since this technique is not well understood and since potentially the  most
significant benefits (i.e., the fastest quenching rates) can be obtained.

III. B.I  Kinetics of NO Decomposition

     The loss of nitric oxide within a gas sampling probe, if indeed  it  occurs,
should be explainable in terms of either hetero- or homogeneous phenomena.  As
reviewed in the TASK I report (Dodge, et. al., 1979), very little information
is available on reaction mechanisms or rates for NO decomposition on walls.
Nevertheless, for walls that are directly water cooled, virtually no NO  can be
lost on either quartz or stainless steel walls for a typical residence time in
a probe of less than one second.  (This statement assumes that the walls are
properly conditioned, as in the case of sampling lines.)  Near the orifice of
a probe tip, wall temperatures may be significantly higher than 300 or 400 °C
and in cases even approach the material softening point.  Under these  cir-
cumstances, the surfaces are certainly catalyt ically active and the potential
for altering flame concentrations exists.  Although a detailed analysis of
molecular diffusion was not performed for this work, it was estimated  that if
the residence time in this portion of the probe is held to less than 10 micro-
seconds, then the molecules undergo only a few collisions with the wall.  Since
the efficiency of NO decomposition should be significantly less than one, then
decompositions of NO by collisions with a hot wall should be negligible.

     Homogeneous mechanisms for the decomposition of NO have been examined by
many workers (e.g., Hanson, et. al., 1974; Flower, et. al . , 1975;  and Koshi and
Asaba, 1979).  A listing of possible reactions which lead to the gas phase
reduction of NOX are listed in Table III-C.  Reactions that oxidize NO to N02
are not included here since no net loss of NO  results.  (Assuming, of course,
that care was taken to avoid loss of N02 in the sampling line and/or sampling
system.)  If the temperature is assumed to be constant, an estimate of the con-
tribution of each of the reactions in Table III-C can then be made.  For small
fractional conversions of NO, the fraction of NO consumed by Reaction  x  is
estimated by
                          fraction =       = k  [x] At                  111-23
                                      N0
                                        0
                       (or = k  [x](M] At where applicable)
                                     111-20

-------
                                     TABLE  III-C
                     REACTION MECHANISM FOR NO DECOMPOSITION
                                                 Rate Constant
                                                              1
Reaction
 Number
 111-15      NO + NO     +  N20 + 0        4.9   x  10
                                                    12
                                    E/R
                                   33,770
 111-16      NO + 0      +  N + 09        2.32  x  109       1    19,445
 111-17      NO + N
 111-18      NO + M
 111-19      NO + H
 111-20      NO +
 111-21      NO + H20
N-, + 0
N + OH
H + HNO
 111-22      NO + H + M  *  HNO + M
1.63 x 10
                                                    13
N + 0 + M     1.41 x
                                                   i21
1.35 x 10
5.75 x 10
OH + HNO      2.0  x 10
              1.8  x 10
                                                    14
                                                    12
                                                    14
                                                    16
-1.5  77,250


      24,760


      28,890


      36,510


        -300
                                                                              Ref.
 Koshi and
 Asabi
 (1979)

Hanson, et.
 al. (1974)

 Baulch,  et.
 al. (1973)

 Baulch,  et.
 al. (1973)

 Flower;  et.
 al. (1975)

 Baulch et.
 al. (1973)

 Baulch,  et.
 al. (1973)

 Jensen and
 Jones  (1978)
1                               "^
 k = ATnexp(-E/RT),  units in cm /mole, sec
                                     111-21

-------
where  [x] represents the concentration of the collision partner and kx,  the
rate constant of Reaction x.  The time increment, At represents an estimated
quenching time, after which the reaction is effectively frozen due to  a  drop  in
the rate constant, concentrations, or both.  Rate constants were obtained  from
literature values and are listed in Table III-C.  Concentrations of stable and
radical species were equated to the maximum equilibrium value for the  three
flat flames examined in this study.  The concentration of NO was assumed to be
5000 ppm, a value larger than the maximum seed concentration used in this  (TASK
II) or the TASK III study.  These concentrations, in terms of mole fraction,
are listed in Table III-D and the fraction of NO lost due to each of these
reactions, as calculated by Equation 111-23, is also given.  The calculations
are based on a (local) probe pressure of 1/2 atmosphere and have been  done for
average static temperatures of 2000 and 1400 K.  In addition, the quench time
was assumed to be one millisecond.  Even by assuming an uncertainty of one
order of magnitude in rate constants or concentrations, it is clear from these
results that only the (overall) three body reaction could possibly contribute
to loss of nitric oxide.  The conclusion that a three body reaction dominates
under conditions of elevated temperature and reduced pressure is extremely
unusual.  It is more likely that the negative temperature dependence used  for
this reaction is too weak and, in fact, the rate constant decreases much more
rapidly at high temperatures than indicated by the rate constants given  in
Table III-C.  Other (overall) three body reactions typically have significantly
higher negative activation energies.  Since this rate constant has only been
measured up to temperatures of 700 K, it should be expected that errors will
result by extrapolation of the rate constant to flame temperatures.   In addition,
Reaction 111-22 should be less important than indicated by Table III-D since
the hydrogen atom concentration will decrease rapidly as the gas is cooled.

     Even if this reaction does cause conversion of NO to HNO, the product will
primarily reform nitric oxide after abstraction of the hydrogen atom,   i.e.,

                                R + HNO - RH + NO                      (111-24)

which occurs at nearly collision frequency.   In this reaction, R represents any
radical species.   Some HNO could possibly be lost via

                               HNO + HNO + N20 + H20                   (111-25)

                               HNO + NO  + N20 + OH                    (111-26)

     or possibly by            HNO + H   * NH  + OH                    (111-27)

     or                        HNO +0   + NH  •»• 02                    (111-28)

     followed by               N20 + R   + RO  + N2                    (111-29)

     and                       NH  + NO  * N2  + OH                    (111-30)
                                     111-22

-------
                                   TABLE III-D

                     ESTIMATED FRACTIONS OF NO DECOMPOSITION
                                                                                O
                                                       Fraction of NO Decomposed
Reaction
Number
111-15
111-16
111-17
111-18
111-19
111-20
111-21
111-22

•'•See text
React ion
Partner
NO
0
N
M
H
H2
H20
H
xM

Estimated Mole Fraction Due to Each Reaction
of Reaction Partner 2000 K
5 x 10~3 3.5 x 10~6
1.5 x 10~5 1.3 x 10~5
7.7 x 10~8 3.8 x 10~6
1.0 8.1 x 10~10
5.7 x 10~5 9.9 x 10~5
2.6 x 10~2 2.4 x 10~4
.15 1.1 x 10~3
5.7 x 10~5 1.0 x 10~2
1.0

1400 K
3.6 x 10~9
2.0 x 10~7
5.5 x 10~6
1.3 x 10"16
7.0 x 10~7
7.0 x 10~7
6.3 x 10~7
2.2 x 10~2


2From Equation 111-23, assuming that pressure = 1/2 atm,  At - 1  millisecond
                                     111-23

-------
 but  none  of  these  reaction mechanisms  is  considered  very  likely.   Reaction
 111-25  is sterically  improbable  and  undoubtedly  represents  a multistep process.
 In any  case,  its reaction rate  is dependent  on the square of the  HNO concen-
 tration which would be quite  low.  Reaction  111-26 is  slow  since  it  has a
 significant  activation energy of 26  kcal/mole (Wilde,  1969).   Reaction 111-27
 is similarly  slow  due to its  high endothermicity  (^24 kcal/mole).   If the
 oxygen  atom  concentration is  high enough  to  allow Reaction  111-28 to proceed at
 a high  rate,  then  it  is highly  likely  that the radical NH will  be oxidized to
 form NO rather  than reduced to  form  No.

      Other reaction mechanisms  (such as to produce HN02) were  also examined,
 but  conclusions similar to those above were  obtained.  No mechanism  could be
 found which  indicated loss of NO  in a water-cooled  sampling probe.   Based on
 the  analysis, the  following quenching  criteria were  selected.

      1.   In  the first portion of the probe tip where high wall  temperatures  (>
          600  K) may exist, the  residence  time of  the gas must  be  less  than 10
          microseconds.

      2.   The  gas sample must  be  "quenched" within 1  millisecond,  and

      3.   The  gas is considered  to be "quenched" when the  total  temperature
          falls  below  1000 K (with an initial temperature  approximately 1800  K).

 This  last criteria is based on  the fact that after 1 millisecond  radicals will
 have  undergone  hundreds of collisions with the walls.  Recombination of these
 radicals,  therefore, must occur within 1 millisecond and equilibrium radical
 concentrations  at  1000 K are  too low to allow any further reactions  with  NO
 before  the gas  is  further cooled to water temperature  (typically  less  than 100
 mi 1liseconds).

 III.B.2   Description of Computer Program  for Probe Analysis

      The  UTRC Probe Design computer program  is based upon an equation  which
 describes  the change in local Mach number as a function of  heat transfer,  skin
 friction, area variations and thermal property changes for  a steady,  one-
 dimensional flow.   Using the  influence coefficient approach of  Shapiro (1953),
 this  equation is

 dM2     9           7 - 1  o  dA           9   dQ     ,     Y ~  1   9      dx
	  (1-M*) = -2(1  + 	 M2) 	 +  (1 + YM2) 	 + YCTU  + 	 M2)  4f 	
 M2                    2       A               cpTs              2           D

                                                                          (111-31)
                           -(1 + YM2) *M- (1 - M2)  H
                                       M             Y
                                     111-24

-------
where M is the Mach number;  YJ  the  ratio  of  specific  heats;  A,  the  cross-
sectional area; Q, the heat  transfer  rate; c  ,  the  heat  capacity  at constant
pressure; Tg, the static  temperature;  f,  the  skin  friction  coefficient;  x,
the axial distance from the  probe orifice; D,  the  tube diameter;  and fl/f>  tne
molecular weight.  The heat  transfer  rate  is  calculated  from

                                dQ = h(Tr  - Tw)  TtDdx                (111-32)

where h is the heat transfer coefficient, T   is  the recovery temperature,  ,
and TW is the wall temperature.  To determine  the wall temperature  distribution
within the probe tip, a computer program  called  TCAL  was  used  (see  end of  this
section).  For both the portion of  the  probe  tip and  the  constant area section
that are cooled directly  by  the water  in  the  coolant  passage,  the wall temperat-
ure is determined by the  overall heat  transfer  rate from  the sample gas  to the
coolant. To provide a conservative  (i.e., low)  estimate  of  the  heat transfer
from the sample to the water, a higher  than  anticipated  coolant temperature
(150 °C) was used in the  calculations.  The  heat transfer coefficients reported
by Kays (1955) for laminar flow and Rohsenow  and Choi (1961)  for  turbulent flow
were used.  For the skin  friction term  (4f),  correlations reported  by Eckert
and Drake (1959) were used;  these correlations  are  applicable  to  fully developed
laminar and turbulent flow in smooth  tubes.

     To calculate the profiles  of temperature,  pressure,  etc.,  the  terms in
Equation 111-31 are evaluated at a  given  axial  location  and  the new Mach number
at the next axial location is calculated  by  a  simple, forward marching proce-
dure.  The gas properties at the new  location may  then be calculated using
standard relationships, such as the ideal gas  law.  When  the flow is supersonic
initially, conditions downstream of a  normal  shock  system are  calculated using
the Rankine-Hugoniot relationships.   The  position of  the  shock  is determined
using one of two procedures.  The position may  be specified  by  the  user  to
occur at a physically realistic location  in  the  probe (e.g., at a bend or sud-
den area increase).  Alternatively, the shock may be  positioned by  the program
so that the stagnation pressure behind  the shock is just  equal  to the desired
back pressure.

     The program performs various checks  on  the  calculations to verify the
stability of the computational procedure.  For  example,  the  program compares
*In continuum flow, the flow velocity is necessarily zero at the wall due to
 frictional forces.  This deceleration of the  flow results  in an increase in
 the flow temperature near the wall.  The flow temperature  at the wall  is termed
 the recovery temperature and is always somewhat  less than  the temperature  that
 would result from decelerating the flow isentropically to  zero velocity  (i.e.,
 the stagnation temperature) (Schlichting,  1960).
                                     111-25

-------
the computed results with results obtained using differential equations  for
selected flow parameters (Shapiro, 1953).  In addition, the step  size  is
selected small enough so that results are essentially independent of the  step
size.  Typically, the initial step size  is between one and two orders  of
magnitude smaller than the orifice diameter.

     In order to estimate the internal wall temperature distribution of  the
probe tip, it is necessary to analyze the heat transfer between the tip  and
the external environment, sample gas, and cooling water.  Heat transfer  calcu-
lations were performed using TCAL, a computer program formulated  at Pratt &
Whitney Aircraft to evaluate the temperature distribution and heat fluxes in
combustors, turbines, and other structural members in gas turbine engines.  In
the program, a finite difference representation of the heat conduction equa-
tion (a time-dependent version of Laplace's equation) is solved by a relaxa-
tion technique.  The inputs to the program include a description  of both  the
geometry and material properties for the hardware being analyzed, the  proper-
ties of fluids around the hardware, and  the heat generated on the surface or
within the device.  Optional boundary conditions include time dependent  surface
and internal heat generation.  Material  properties are permitted  to vary with
temperature and boundary temperatures are allowed to vary with time.

III.B.2a.  Sudden Expansion Losses

     The computer program (probe design  deck) used in this study  for analysis
and design of gas sampling probes assumes that the gas flow is one-dimensional.
It, furthermore, requires that the flow  within the probe be attached to  the
probe walls; flow separation is not modeled.  In the case of probes designed to
operate with a choked orifice followed by a supersonic expansion, this require-
ment is usually satisfied.  If this type of probe is operated unchoked, then it
is probable that the subsonic flow within the expanding probe tip will separate
from the probe walls due to the high divergence angle of these walls.  Detailed
analysis of the boundary layer accompanying subsonic flow in the  tip of one the
large probes used in this study revealed that the subsonic flow is indeed
separated.  Furthermore, it was demonstrated experimentally that  for probe
operation at high back pressure (>l/2 external pressure) the probe orifice was
unchoked.   Consequently, the following procedure was included to  account  for
the pressure loss associated with separated internal flow.

     During the process of diffusing a subsonic flow, frictional  forces  and the
adverse pressure gradient may be sufficient to dissipate the forward momentum
of the flow, i.e., the flow will separate.  Practical subsonic diffusers  limit
the wall angle of the device to a few degrees.  Wall angles greater than  this
limit (such as encountered with flows over rearward facing steps  or with  flows
in probes with short, high area ratio tips) tend to cause flow separation.  The
loss in stagnation pressure due to flow  separation is termed the  sudden  expan-
sion loss .
                                     111-26

-------
     The  sudden expansion loss  may be calculated by assuming that the static
 pressure  in  the separated region is equal to the (unknown) static pressure at
 station 2  (see  Fig.  III-l).   Thus, for one-dimensional flow, the continuity and
 momentum  equations may  be applied  to the control volume shown in Fig. III-l
 (dashed line).   After some manipulation, the system of equations becomes:
                    PT      \PT/1    \PT/1  \A1   /                  (111-33)
                    PT
                     Ll
                             PT2    (A/A*)9
                            _ f  _ _ *                           (111-34)
                             PT1    A2  (A/A*)L
                   <
where  f = P(l + "VM )  and  represents  the  stream  thrust  per  unit  area.   A  is
the  flow area, M,  the Mach  number, P,  the  static  pressure, P™,  the  stagnation
pressure, and *y, the  ratio  of  specific heats.   A   is  the area at  the  throat
and  the subscripts 1  and  2  refer  to  the  upstream  and  downstream conditions,
respectively.  Writing Equations  111-33  and  34  in terms of the  Mach numbers,
Mi and ^2, and the area ratio, ^l f^i ,  and  solving the  equations simultaneously
yields an equation which  may be solved explicitly for  M2 .  Then,  the  pressure
ratio, Pj2/Pxi' may be calculated from either Equation 111-33 or  34.   The sudden
expansion loss is  1 - Pf2/PTl •  ^t *s> °^  course,  assumed  that  the  initial con-
ditions (M, , Pj) and the  points of flow  separation and reattachment (which
defines A2/Ai) are  known  to sufficient accuracy.

III.B.2b.   Aerodynamic Quench

     From extensive use of  the computer  program,  some  qualitative features on
the operation and  design  of 'aeroquench' probes have been obtained.   These
facts are reviewed  here along with a more  detailed description  of the  phenomena
of aerodynamic cooling.
                                     111-27

-------
                                MODEL FOR CALCULATING SUDDEN EXPANSION LOSS
i
r-o
00
   01
   Ol
   I

-------
     An aerodynamic quench  in  a  sampling  probe  is  accomplished  by reducing the
static temperature to  some  level  determined  from the  kinetic  rates and maintain-
ing it at this  level until  the stagnation temperature of  the  gas  is also reduced
to near this  level.  The  reduction  of  the static temperature  is achieved by
accelerating  the  flow  to  a  high  supersonic Mach  number.   During this process,
the random kinetic energy in the  gas  is converted  to  directed translational
energy and, hence, a net  cooling  of gas occurs.   The  reduction  of the stagnation
temperature is  achieved via heat  transfer to cooled walls of  the  probe.   (The
stagnation temperature  is the  temperature the gas  will  reach  if decelerated
adiabatically to  zero  velocity.)  This heat  transfer  to the walls is of  prime
importance since  without  it the  gas would eventually  shock heat during its
recovery to subsonic flow conditions  and  return  to temperatures approaching the
original flame  conditions.  Since an  aerodynamic quench is obtained by accelerat-
ing the flow, a significant stagnation pressure  loss  results  due  to friction.
Therefore, whether an  aerodynamic quench  can be  achieved  is highly dependent
upon the pumping  system used with the  probe.  The  pumping system  sets the  back
pressure which  is the  maximum  pressure behind the  shock system  that terminates
the supersonic  portion of the  flow.

     In a probe designed  to quench  aerodynamically, the flow  is accelerated
from a Mach number of  unity at the  orifice to a  high  supersonic Mach number via
the area expansion in  the probe  tip.   This expansion  must be  large  enough  to
achieve the reduction  in  static  temperature  required  yet  not  so large as to
cause friction  to reduce  the stagnation pressure below  that achievable by  the
pumping system.   Larger area ratios would increase the Mach number  and con-
sequently increase frictional  losses.  The flow  then  enters the constant area
section where the stagnation temperature  is  reduced by heat transfer.  Some
heat transfer also occurs in the  tip but  is  generally only a  small  part  of the
total heat transfer because the  residence time of  the flow is short  in this
region.  In the constant  area  portion, both  the  stagnation pressure and  the
supersonic Mach number  of the  flow  are reduced due to friction.   If calcula-
tions indicate  that the Mach number is reduced to  unity in the  constant  area
section, either a shock must instead occur at some upstream location where the
shock may be  stabilized or  the flow in the probe is subsonic  throughout.   In
either case, no aerodynamic quench  is  possible.   The  constant area  section must
terminate prior to the  occurrence of choking but only after the desired  reduction
in stagnation temperature is achieved.  This  section  is generally terminated by
a sudden expansion to  stabilize  a shock system that reduces the Mach number to
a subsonic value.  The  expansion must be  large enough to  result in  a relatively
low Mach number so that additional  stagnation pressure  losses are small; other-
wise,  the flow will choke in the  subsonic portion  of  the  probe.   It  is  impor-
tant to note that the  probe must  not contain a bend prior to  the  desired
shock system  location.
                                     111-29

-------
III.C  Design of Probes

     Using the above assumptions  for the probe model and experimental  conditions,
two sets of  probes were designed.  One set was designed for sampling over  the
flat flame burner, while the other was designed for sampling exhaust from  the
IFRF burner  or the FT12 combustor.  In these design studies, the primary design
Criteria was to reduce the gas temperature below 1000 K within one millisecond
Csee Section III.B.I.a).  Furthermore, it was assumed that the back pressure
was 0.35 atmospheres (266 torr),  a pressure that should be easily achievable
for most sampling systems.  A special effort was made to develop a probe design
for which true aerodynamic quenching would be expected.

III.C.I  Probes for Combustor Measurements

     For sampling the exhaust of  the IFRF burner and the FT12 combustor, two
water-cooled probes were designed:  one for cooling by expansion and the other
quenched by  convection.  These probes are denoted the reference probe  and  EPA
probe respectively and drawings of these probes are depicted in Fig. III-2.
•As apparent  from this figure, pressure taps and thermocouples have been located
in these probes to facilitate comparison of experimental and computed  data.
Externally,  both probes are identical and photographs of the reference probe
with its water-cooled mount are shown in Figs. III-3 and III-4.  Due to their
size, they have been denoted macroprobes.  They were each designed to  enter
from a side window of the combustor but with a right angle bend to face the
oncoming flow.  The probes are sufficiently long so that they could traverse
across the internal dimension of  the combustor along the optical path.

     The geometry of the reference probe with two sections of constant area
follows the guidelines described  earlier (Section III.B.2b) for a probe that
quenches by expansion.  As will be apparent, the second expansion has been
located far enough downstream so  that the recovery shock returns the gas to
temperatures near 1000 K when the initial gas temperature is 1800 K.  Also
observe that the tip has been carefully contoured to minimize the possibility
of coalescence of compression waves (i.e., shock waves) by providing a smooth
transition at the junction of the tip and the first constant area section.   It
is very important to note that even with this ideal design, aerodynamic quench-
ing could not be achieved (based  on computer program prediction) unless the
back pressure was reduced to below 7 percent («^50 torr) of the ambient flame
pressure which in this case was atmospheric pressure.  This pressure was much
lower than the desired operating  pressure of 0.35 atmospheres and, as described
in the experimental section, creates some special sampling problems.

     A calculated profile showing the static temperature and pressure  and  the
total temperature vs. time is shown in Fig. III-5 for assumed flame tempera-
tures of 1800 and 1400 K.  In the case of the higher temperature, internal
positions from the probe tip are  also provided.  Although these curves indi-
cate that the static temperature  rapidly decreases to below 1000 K (in less
                                     111-30

-------
                                             DRAWINGS OF MACROPROBES

-------
                                                  REFERENCE PROBE

-------
TIP OF REFERENCE PROBE
                                           FIG. 111-4
                                        79-10-85-?

-------
                      CALCULATED TEMPERATURE AND PRESSURE PROFILES FOR REFERENCE PROBE
       1800
I
UJ
    QC
    OC
    LJJ
    CL

    5
    LLJ
       1600
       1400
       1200
1000
        800
        600
                 23
                                                  ---- Tex, = 1400K

                                              POSITION FROM TIP (CM) AT 1800 K

                                        7   10    15    2022      25        30
                                                        TT      T
                                                          SUDDEN
                                                         EXPANSION
                                                                                                            0.10
                                                                                                            0.08
                                                                                                     0.06 <

                                                                                                         LU
                                                                                                         QC
                                                                                                            0.04
                                                                                                      0.02
 <0
 I
 01
 I
        400
                                                         J	I    I   I  I  I  I I
                  10'5
                                         10'4
                                                      RESIDENCE TIME (SEC)
10'3
10-2
                                                                                                           O

-------
than 10 microseconds),  the  total  temperature is  still high and does not fall
below  this  level  until  about  1/2  millisecond for 1800 K inlet temperatures.
The two primary  features  in the  static  temperature curves  are the minimum at
about  25 microseconds  indicating  the  termination of the divergence in the tip
and the rapid  rise  due  to shock  recovery  at  about 200 microseconds.

     Calculated  cooling curves  for  the  EPA probe which cools  primarily by
convection  are shown  in Fig.  III-6  for  the two  assumed inlet  temperatures,  1400
and 1800 K.  For  these  profiles  a back  pressure  in the probe  was  assumed  to  be
0.35 atmospheres  (266  torr).  According to these calculated curves, the static
temperature  (nearly equivalent  to total temperature for subsonic  flow) reduces
to below 1000 K  in  approximately  0.3  milliseconds.   For comparison, calculated
data from the  reference probe when  operated  at  the same back  pressure is  pre-
sented.  In  this  case,  the  reference  probe cools by convection.   This probe
which  was designed  to  achieve an  aerodynamic quench is clearly relatively poor
at cooling  by  convection  since  the  static  temperature does not reach 1000 K
until  2.3 milliseconds, a difference  of nearly one order of magnitude from the
EPA probe.   The  inflection  near  this  location is due to the area  change at 22
centimeters  from  the  tip.

     For both  probes,  when  cooling  by convection,  it may be noted  that  the
program indicates the  existence  of  a  normal  shock within a centimeter of  the
tip orifice.   In  each  case,  temperature recovery occurs within several  micro-
seconds and  returns the gas  from  less than 1000  K to within approximately 100 K
of the original  gas temperature.  It  is this operating mode that  is  probably
typical of most  probes  previously believed to be aerodynamically  quenched.  The
probe  is indeed  choked, the  gas  flows supersonically for a period,  and  the
initial quenching rates are  rapid with  temperatures  falling below  1000  K;
however, a  shock  system quickly  forces  the gas to the  elevated  temperatures and
the initial  quenching,  although  real, is only temporary.   Further  examination
of the EPA probe  design with  the  computer  program indicated that no back
pressure existed  for which  the shock  system  could be pulled into  the  constant
area section and  produce  an  aerodynamic quench.   The difficulty was  not due to
the small discontinuity in  the slope  at the  end  of  the  tip but  rather due to
excessive friction  and  pressure  losses  in  the probe.

III.C.2  Probes for the Flat  Flame  Burner

     Several probes were  selected for sampling over an atmospheric  CH^/02/N2
flat flame.  The  orifice diameter for these  probes  (miniprobes) was selected  to
be 0.025 inches  (635 microns) which was large relative  to  the  quartz  microprobes
used in the  first phase of  this  program but  was  necessary  to  supply the mass
flow required by  the analytical  instrumentation.   Using this  orifice  diameter
and assumed  flame conditions  of  2000  K  and one atmosphere, no  practical geometry
could be found over an  experimentally realizable back  pressure range  for  which
an aerodynamic quench,  as defined above, could be achieved.   Consequently, all
design efforts for  these probes were  focused at  convection cooling below  1000 K
and within one millisecond.
                                     111-35

-------
                                        CALCULATED COOLING CURVES FOR MACROPROBES
                                 0.91
i
u;
      o
      I
      O
      I
      IX
      I
                        1800
                        1600  -
                     5  1400 —
                     01
                     tr
                     I
                        1200  -
                     O
                     C/3
                        1000
                        800 —
                        600 —
                                    1.0
        POSITION (cm) FOR TT = 1800K AND REFERENCE PROBE
                        END OF TIP                AREA CHANGE
                2.0      3.0 i  4.0 5.0        10    15  201  25    30
                                                                                   BACK PRESSURE = 0 35 aim
                                                                                         EPA PROBE
                                                                                  	REFERENCE PROBE
                                     T = 1400K    \ TT =
                                   TT = 1800K \
TO
10'4                    ID'3
  RESIDENCE TIME (sec)
                                                                                                                     T]
                                                                                                                     O

-------
     Designs of  two water-cooled probes were  selected  for  cooling  by  convection.
One probe was entirely  stainless-steel.   The  computer  calculations, however,
indicated that the wall temperature  of  this probe was  reduced  to only about  750
K at the location for which  the gas  residence  time was  approximately  10  micro-
seconds.  This temperature was not considered  low enough to minimize  wall
reactions in the probe  tip.   In order to  satisfy this  quenching criterion, a
copper tipped, water-cooled  probe but which is  otherwise identical  to the
stainless steel  probe was also constructed.   Calculations  for  the  copper-tipped
probe indicated  not only that the wall  temperature fell to around  500 K  within
a gas residence  time of 10 psec but  also  that  the  initial  cooling  rate of  the  gas
was significantly faster.  A relatively blunt  tip does  offer an alternative  to
the change in material, but  this was not  considered  feasible since  such  a  probe
could significantly perturb  the flame and  since flow separation at  the entrance
of the tip (due  to the  large  angle)  was considered to  be very  likely.  A
catalytic surface effect due  to the  copper was  considered  possible  but believed
to be negligible because of  the short residence time in the tip.  A drawing and
photograph of the stainless-steel probe are shown in Figs. III-7 and  III-8
respectively.  The third, water-cooled  probe was constructed completely  of
quartz.  Its design was similar to the  quartz microprobe used  in Task  I  of this
study (Dodge, et al. , 1979),  however, the  orifice was  enlarged to 635  microns
by shortening the tip (see,  for example,  the description by Fristrom  and
Westenberg (1965)).  For comparison, an uncooled stainless-steel probe was
constructed which, except for a smaller outside diameter (0.635 cm) and  no
cooling passages, was geometrically  identical to the other metallic probes.

     Model predictions  for the copper tipped and stainless-steel tipped  probes
are compared in  Fig. III-9.   The positions apply to  the stainless-steel  probe
although they also approximate the positions  for the copper tipped  probe.
Clearly, there is little difference  between the calculated cooling  curves  for
these probes after approximately one millisecond of  residence  time; however,
the initial cooling rates are sufficiently faster with  the copper tip  so that
at residence times less than  200 microseconds,  gas temperatures are nearly 150 K
cooler.  Note that for  both  of these probes, the gas flow chokes at the  tip,
accelerates supersonically and then  shock heats within  one centimeter  of the
probe orifice.

     The effect  of back pressure on  temperature profiles and cooling  rates
was also examined.  Model predictions for  the  copper-tipped probe  at  two back
pressures, 0.35 and 0.167 atm  (266 and 127  torr, respectively), are  shown in
Figure 111-10.   It is clear  that at  the lower back pressure the recovery shock
occurs later in  time and at  a cooler static temperature, the static temperature
returns to temperatures 100  K lower  than  those  at the  higher back  pressure, and
the total temperature (nearly the same  as  the static temperature under subsonic
flow) reduces to below  1000 K in about  140 microseconds, approximately half  the
time required at the higher pressure.   In addition,  a  50% drop in  the  back
pressure provides another advantage  to quenching since  the rates of bimolecular
reactions are dropped by a factor of four, and  even  unimolecular reaction  rates
slow by a factor of two.
                                      111-37

-------
                                                   DRAWING OF MINIPROBE
                                          TIP
I
LJ
CD
                                                                                        XX  XX  X  X
                                                                                                          0.95cm
    I
    r
                                                                                                                TJ

-------
                                           STAINLESS STEEL TIPPED MINIPROBE
                          GAS SAMPLE OUT
                            WATER IN AND OUT
                                                    f .^  -..__, ... ' _^g,__^ -jy- ' ;	
                                                           -J .  I - -J   ,
                                                     cm 1  2 3
                                                                                               ORIFICE. 635 MICRONS -
ID
I

s
I
U


-------
                                       CALCULATED COOLING CURVES FOR MINIPROBES
          2000
i
i~
o
     cr
     ID
     cr
     LU
     Q.
     O
 o
 I
 CO
 en
 I
           1600
1200
           800
           400
                            INITIAL STAGNATION TEMPERATURE = 2000K
                            A',Si IMI D RA( K PRESSURE = 0 35 aim
                                                            WATER-COOLED
                                                            6,15 MICRON ORIFICE. ML TALI 1C PROBES
                    0 70
                                        POSITION (CM) FOR S.S TIPPED PROBE
                                                 2              34
      7  8  9 10  12   15  20
                                                                          T
                                                                               IIIITT
                T
T
                                                                  END OF
                                                                 PROBE TIP
                                                                                   	Cu TIPPED PROBE

                                                                                   	 SS TIPPED PROBL
                          10'
                            ,-5
                                                   ,-4
                                                 10'
                                                RESIDENCE TIME (SEC)
10'
  ,-3
    10'
      -2
                                                                                                                      O
                                                                                                                                CD

-------
                   CALCULATED COOLING CURVES FOR MINIPROBES AT VARYING BACK PRESSURE
            1400 -
                                                                             Cu-TIPPED PROBE


                                                                             TTOTAL = 1800K
                          10'
CO



o


00
U1
6   8 10'4

RESIDENCE TIME (SEC)


-------
     Detailed calculations were not performed for the quartz  probe  or  the
uncooled probe.  The primary differences for these probes will of course be
surface material for the former and, for the latter, surface  material  at elevated
temperatures and lack of rapid cooling (some cooling is expected since the
probe itself will radiatively and conductively cool and this  cooling in turn is
passed onto the gas via conduction).

III.C.2a  Microprobe

     In the course of analyzing the behavior of probes, quartz microprobes were
also examined since they had been used in the first part of this program.
These calculations using the current version of the UTRC Probe Analysis computer
program indicate that supersonic flow cannot be sustained within the tip region
of so-called "microprobes".  Since supersonic flow can be maintained in larger
probes (if the back pressure is set at a prescribed level), it is desirable  to
explain the differences in the flow characteristics within each type of probe
in terms of the physical dimensions of the probes.

     For a constant property flow, it can be shown that the local Mach  number
varies according to:

         dM2    1 - M2         dA          0  dT      0    dx
                                                .
                          = -2 	 + (1 + YM2) 	° + YM* 4f —        (111-35)
         M2  j + Y - 1 M2      A              TQ
where M is the local Mach number, A, the area, Y, the ratio of specific heats;
T , the stagnation temperature;  f, the skin friction coefficient, x, the distance
from the probe orifice, and D, the internal diameter.

     Since this relationship holds for a flow passage of any size, the difference
between flows within small and large probes must lie in the relative contribution
of each term (area change, heat  transfer, friction).  For a supersonic flow  (M
> 1) within a passage of increasing area (dA > 0) and with cooling (dT  < 0),
the Mach number will increase  if the absolute value of the area change and heat
transfer terms in Eq. (111-35) exceed the frictional contribution; the Mach
number will decrease if the opposite is true.

     For example, consider the following case for both a micro-and macroprobe.
Assume that the probes are used  to sample a flow whose stagnation pressure and
temperature are one atmosphere and 2000 K, respectively.  Assume that the flow
is choked at the entrance of a probe whose tip is a conical section with a
half-angle of 7 degrees.

     For a microprobe, assume  that the initial probe sample passage diameter
is 0.01 cm (4 mils).   It can be  shown that the flow accelerates initially (even
in the absence of heat transfer).  The Mach number reaches a peak value of 1.71
                                      111-42

-------
at 0.015 cm from the tip and thereafter  decelerates  to  unity  (chokes)  at  0.167
cm.  This second choke point is  unrealistic  and  indicates  that  instead a  shock
must occur  near the tip or that  subsonic  flow must  exist  throughout.   The
reversal of the Mach number change  is  explained  by noting  that  the  flow in  this
case is laminar.   It can be shown for  such a flow  that  as  the distance from the
orifice increases, both the area  change  and  heat transfer  terms  decrease  and
the frictional term remains approximately  constant.   Thus,  the  friction term
eventually dominates the change  in  the Mach  number equation.

     For a macroprobe, assume  that  the initial diameter is  0.2  cm (0.079  in.)
which is identical to the orifice of the macroprobes  in this  study.  In this
case, the flow within the probe  is  turbulent.  It  can be shown  for  fully
developed turbulent flow that  all contributions  to the  change in Mach  number
decreases as the diameter of the  passage increases.   Calculations indicate  that
friction effects eventually dominate (as they do in  the laminar  case)  and can
choke the flow under the right circumstances; however,  the  distance to the
choke point may be long enough to provide  sufficient  length to  achieve an
aerodynamic quench.  Choking is  avoided  by encouraging  the  formation of a
normal shock (such as by a sudden expansion)  and slowing to a low subsonic Mach
number.  Consequently, the rate  of  frictional losses  is drastically reduced.

     Thus, in either case, the friction  contribution  eventually  dominates the
variation in Mach  number.  If  supersonic flow exists  within a microprobe, it
cannot be sustained for any useful  length; rather  a  normal  shock must  exist
very near the probe tip and a  low subsonic Mach  number  must exist thereafter.
This unrealistic shock location  (within  a  few orifice diameters  of  the tip)
together with the  fact that boundary layer growth has been  ignored, lead  to  the
conclusion that supersonic flow  is  not likely within  a  microprobe.  Boundary
layer growth is less important in a macroprobe,  and  the required shock location
is more realistic.  Thus, supersonic flow  can exist within  a macroprobe.
                                      111-43

-------
                         IV.  EXPERIMENTAL  RESULTS
     The objectives of  the  experimental measurements  on  the  three  combustors
were two fold.  First of  all, probe measurements  of  seeded nitric  oxide  would
be made, compared with  expected  concentrations, and  causes for  any existing
discrepancies between the various  probe measurements  and/or  the  expected
concentrations analyzed.  Secondly, sufficient  probe  data (specifically,
concentration and temperature profiles) must be obtained so  that subsequent
optical measurements can  be properly  interpreted.  These measurements  along
with experimental data  on probe  behavior  are presented in this  chapter.
IV.A.  Flat Flame Burner

     Three flame stoichiometries,  4> = 0.8,  1.0, and  1.2, were examined and the
run conditions are listed  in Table II~A.  With  the nitrogen purge passing
through the optical ports, vertical and horizontal temperature profiles were
obtained.  A typical horizontal  profile (for  4>  = 0.8) is shown in Fig. IV-1 .
These measurements are corrected  for radiation  losses as outlined in Task I
Report by Dodge, et al. (1979) and were obtained along the optical axis.  The
burner surface was located 2 centimeters below  this  axis while the visible
flame sheet was a few millimeters  above the burner surface.  Estimates of
uncertainties in the radiation corrections  and  individual and repeated measure-
ments are indicated in the error  bars.  Vertical profiles of uncorrected
thermocouple temperatures  for the  three flames  are shown in Fig. IV-2.  Since
the optical beam is less than a  centimeter  in diameter and centered at 2
centimeters above the burner, the  beam encompasses a region for which devia-
tions (due to height) are  less than + 15K.  As  may be observed by comparing
these two figures, radiation corrections are  on the  order of 140 K for these
f1ame s.

     Measurements of stable species using the Scott  Instrument package and
the quartz, water-cooled probe (all water-cooled probes produced the same
results  within 10 percent) are reproduced in Table IV-A.  Equilibrium data at
the adiabatic flame temperature  (Gordon and McBride, 1971) for the flat flames
are also presented in this table  for comparison.  It may be noted that reasonable
agreement is obtained between the measured  and  equilibrium values.  Noticeable
differences are observed for the  CO and CC^ concentrations.  Although these
differences may be due to  interconversion within the probe, most of these
differences may be explained by differences in  actual and adiabatic flame
temperatures and to incomplete combustion of carbon monoxide.  Only in the case
of the stoichiometric case ( 4> =  1.0) is the measured total of the CO and C02
concentration significantly different from  the  equilibrium calculations (^8%).
(Data on unburned hydrocarbons are not reported since their concentrations even
in the fuel-rich flame were less  than 1000  ppm.)
                                      IV-1

-------
                          HORIZONTAL TEMPERATURE PROFILE OVER CH4/O2/N2 FLAT FLAME
                                                        = 08
                                              O THERMOCOUPLE TEMPERATURE
                                                CORRECTED FOR RADIATION
                                                    -2000-
I
K)
     -11
-10
-9    -8
o
-5
-4
-3-2-101    23


 POSITION FROM CENTERLINE (CM)
10   11
                                                                                                        31

-------
                                                               FIG. ET-2
VERTICAL TEMPERATURE PROFILE OVER CH4/02/N2 FLAT FLAME
                          O 0 = 0.8

                          A  = 1.0

                          D  = 1.2
        1700
        1600
     cc
     LU
     a.
        1500
        1400
                       UNCORRECTED FOR RADIATION
                        I
                        1            2

                     HEIGHT ABOVE BURNER (CM)
                                                              79-10-85-14
                            IV-3

-------
                           TABLE IV - A
                MOLE PERCENT OF STABLE SPECIES FOR
                       THE FLAT FLAME BURNER
                            (Wet Basis)
                           Experimental

                  CO1     CO,1     H~02      N03     Temp.  (K)
0.8
1.0
1.2
3.2
0.24
—
.01374
.064
4.1
6.6
6.55
4.8
12.6
14.1
17.8
77.6
78.7
73.3
1740
1815
1800
Equilibrium
*
0.8
1.0
1.2
o2
3.15
0.13
0 . 7ppm
CO
0.0051
0.109
3.60
co2
6.43
7.09
5.50
H20
12.9
14.3
15.7
N2
77.4
78.1
72.6
Temp. (K)
1765
1905
1904
1.  Measured values but corrected for the presence of water vapor.
2.  Water estimated from known input conditions.
3.  Nitrogen calculated by difference.
4.  Error - 40% of value.
5.  Based on equilibrium flame temperature.

Except where noted, the uncertainity in the experimental concentrations
is approximately ± 5% of reported (experimental) value.
                                IV-4

-------
     Measured  concentrations  of nitric oxide were on the order of 5, 30, and 30
 ppm  for  the  three  flames,  respectively.   These values at flame temperatures and
 for  the  given  optical  path length are much smaller than that necessary to
 produce  a  reasonable  signal-to-noise ratio for the optical measurements.
 Typically, concentrations  of  700-1500 ppm are required for the UV resonant
 lamp measurements  and  even higher NO densities are required for the infrared
 gas  correlation  technique.   Consequently,  all subsequent studies  were made
 using nitric oxide premixed with the inlet gases.

     Using the stainless  steel  tipped, water-cooled probe and the TECO CLA,
 nitric oxide horizontal profiles were obtained for a given seed level.  Profiles,
 normalized to  the  input seed  level for the flames  4> = 0.8 and 41 = 1.2 are shown
 in Fig.  IV-3.  Data for the 4> = 1 .0 flame  is nearly identical to  the data for
 the  $ =  0.8  flame  except  the  centerline  fraction is slightly higher (^0.88)
 for  an NO  seed level  of 840 and 980 ppm calculated on a wet and dry basis,
 respectively.  For the lean flame, NOX values were typically 5 to 7 percent
 higher than  the  NO and, for the stoichiometric and rich flames,  they were about
 3 percent higher than  measured  NO.  The  excellent  repeatability of the burner
 and  sampling conditions is indicated by  the double set of points  on the right-
 hand side  of this  figure.

     Vertical  profiles for the  three flames and the three water-cooled probes
 are  reproduced in  Fig. IV-4.   (Note, the data in this figure are  not normalized.)
 Agreement  between  these profiles is generally quite good (within  6%) except for
 relatively low values  obtained  by the stainless steel tipped probe when sampling
 the  rich flame.  Since the front part of this tip  (J" 1 cm)  becomes very hot
 (with a  red-orange glow),  it  is believed that catalytic reactions take place
 similar  to those occurring in an N02 to  NO converter.   This conclusion is
 consistent with  the computed  results from  TCAL indicating that the stainless
 steel tip  is insufficiently cooled.   Although residence times in  this  portion
 of the probe are very  short (« 1 msec),  the wall  temperatures very near  the
 orifice  are  significantly  hotter than in a stainless steel  converter (1100  to
 1200 K vs. 1000  K).  The  profile may be  associated with the presence of hydrogen
 that would be  expected to  decrease with  height above the flame.   Although this
 mechanism was  not  verified,  it  seems highly likely considering the strong NOX
 reducing effect  that hydrogen has in hot  stainless steel tubes (Benson and
 Samuelsen, 1976, 1977).   The  relatively  low values obtained in the  = 0.8
 flame for the  copper-tipped probe are unexplained.  Although this difference of
 7 percent may  be due to a  catalytic  effect of a copper surface, it is  unclear
 why good agreement  is  found for the  other  flames.   In any case, this difference
 is considered  to be small.

     From the  above results,  it is seen  that the recovery of the  nitric oxide
 seeded into  the  flame  is not  quantitative.   At the seed concentrations,  approxi-
mately 15, 12, and  38  percent of the initial NO was lost for the   = 0.8, 1.0,
 and 1.2  flames,  respectively.   Attention was directed  to determining whether
 the loss occurred  in the  inlet  gas lines  (for the  unburned/premixed gas), in
 the flame itself,  or the  probe/sampling  system.   Losses in  the post-flame zone
were believed  to be unlikely  since essentially flat vertical profiles  were
obtained (Fig. IV-4).
                                      IV-5

-------
                               NORMALIZED NITRIC OXIDE PROFILES OVER CH4/O2/N2/NO FLAT FLAME


                                                           c6 = oe AND  0=12
                                             •


                                             O
                                                    NO SEED LEVEL

                                                   'WET'     'DRV


                                                 08  RSO       971 PPM

                                                 12  H,'H      1011 PPM
                                                               -1.0
I
o>
                                                                 -A
                                                               0.8
                                                     •*•
                                                      O
                        8    8    8
                                             "O --- O
                                     -- 06$-
                                     	6	Q -
                                               I
                              I
                          I
                                       Q
                                       LU
                                       UU
                                       CO

                                       O
                                       z

                                       Q
                                       LU
                                       QC


                                       CO

                                       LU
I
                                                               0.4
                                                               0.2
I
1
1
          -11    -10
-9
     o
     I
     CO
     01
     I
-7    -6    -5-4-3-2-101    234


                  POSITION FROM CENTER OF BURNER (cm)
                                                      10   11

                                                                                                                           co

-------
                                                                FIG. ET-4
VERTICAL PROFILES OF NITRIC OXIDE OVER FLAT FLAME BURNER

                 O  s.s. TIPPED)
                 A  Cu TIPPED > WATER COOLED PROBES
                 D  QUARTZ  1
900
1
LU
2
O.
| 800 j
0
S
DC
Z
LU
Z 700
0
O
O
D
LU
DC
| 600
LU
1
= 0.8
SEED 971 PPM


j- g B D B
A
A

~


_
I I
2 3
j
.# = 1 .0
SEED 980 PPM
. A A A A
8 § e 8 8
r







j


—


—
I I
2 3
HEIGHT ABOVE BURNER (CM)
#=1.2
900
P
LU
a.
CL
w 800
0
f—
DC
1—
Z
LU
O
8 ?0°
o
Z
a
1 1 I
LLJ
DC
g 600.
LU t
SEED 1011 PPM


—


) s g 8 0
e e 8
_ 8 o ° °
i i








               1           2          3
                   HEIGHT ABOVE BURNER (CM)


                             IV-7
79-10-85-11

-------
     These loss figures can be slightly adjusted to account  for  phenomena
associated with the flame and experimental apparatus.  First  of  all,  some NC^
was present as indicated by NO  readings, and, consequently,  the percent
losses can be decreased by 3 to 5 percent.  Although  it  is uncertain  whether
the presence of NC>2 is due to the flame or probe, the  fraction of  NC^ was not
large enough relative to experimental uncertainties to warrant a detailed
investigation.  Alternatively, if one assumes that the NO formed in the  flame
without  seed NO  is  also formed when NO  is added, then the  losses in  the
stoichiometric and rich cases can be increased by approximately  3  percent
(about 30 ppm in 1000).  The estimated losses at centerline of NO  may be
decreased by about 3 to 5 percent due to dilution from the nitrogen purge in
the optical ports (see Dodge, et. al., 1979).  Uncertainties  also  include the
inaccuracies in blending from the mixing apparatus and in the calibration and
analysis.  It is estimated that the sum of these uncertainties is  on  the  order
of 3 percent since measured NO concentrations generally  agreed to  within  3
percent of the calculated values when NO was blended  only with nitrogen,  and
the gas sample was extracted within 1 mm of the burner surface with no flame
present and with the nitrogen purgp off.  This procedure provides  an  extremely
powerful and important experimental tool to check the operation  of the complete
apparatus.  It was used to verify not only the accuracy  of the blending methods
but also the integrity and behavior of the sampling system.   For example,  this
procedure was used several times to identify the existence, although  generally
not the location, of sample line leaks and indicated  the necessity for condi-
tioning new stainless steel lines used in the sampling line or detector system
(see Section III.A.3).  This technique is particularly advantageous over  other
alternative tests such as a vacuum check for leaks, since it  can easily be
performed prior to and after any given experiment and does not require the
opening and resealing of the sample line or probe connection  to  the sample
line.   (Moreover, it also identifies the existence or lack thereof of other
loss mechanisms of NO such as line conditioning or other line losses.)

     With all of the adjustments and uncertainties mentioned  above, the percent
NOX lost for these flames becomes 9, 8 and 34 percent respectively with an
uncertainty of about ± 5 percent in each number, i.e., for the lean flame, the
estimated real loss can range from 4 to 14 percent.   If  the natural NO formed
in the flame is not included, then these numbers become  9, 5, and  31  percent
for the 4> = 0.8,  1.0,  and 1.2 flames, respectively.

     To check whether NO was conserved prior to the exit from the  surface of
the burner, the following test was made.  The probe tip  was placed near the
surface (within 0.4 cm) but still above the flame sheet  and NO and NO mea-
surements were made.  Then the mass flow of the carrier  gas (N^) was  increased
to push the visible flame sheet well above the probe  tip ( ^ 2 cm)  and NO
and NOX measurements were repeated.  This test was performed  using the copper
tipped probe on the lean and the rich flames with approximately  1000  ppm  of
                                      IV-8

-------
seed NO and no  nitrogen  purge.   With  the  flame below the probe tip for the lean
flame, the probe measurement  indicated  a  9  percent  loss of NO .   With the flame
pushed above  the tip  and  correcting  for increased diluent in the flame,  measure-
ments of NOX  were  2 percent higher  than the calculated value, with 10 percent
of this NOX measured  as N02.   In the  case of the NOX value,  exact correc-
tions due to  changes  in  the viscosity and quenching efficiency of the carrier
gas (Dodge, et. al.,  1979a) were not  made since the constituency of the  carrier
was uncertain after it passed  through the converter.   Estimates  for both NO and
N0x, indicated  that less  than  a  3 percent correction would be required.   For
the rich flame, N0x measurements indicated  a loss of 31 percent  of the seed
NO with the probe  at  0.4  cm above the burner;  but with the flame pushed  above
the probe tip,  the  loss was only 6  percent  of  the original NO and at  least half
this amount could  be  due  to viscosity and quenching effects.   The remaining
differences are considered to  be negligible.   Again,  10 percent  of the measured
NOX was N02 .  (in  the rich case, the  NO  converter  (molybdenum)  could be
operated for  a  short  period of  time,  approximately  one minute or so,  before
reconditioning  was  necessary).

     These results  indicate that NO   is conserved in the gas  handling apparatus
                                    A.
and up to the burner.  One may  argue  that by lifting the flame off the burner,
the temperature in the sintered  copper  surface is reduced and consequently
catalytic action which would  depend  on  stoichiometry is reduced.   It  is  not
believed that this  is the case,  however,  due to other evidence obtained  at
UTRC, (discussed later in this  section).

     These measurements  also  verify  the operation of  the sampling line and
analysis system.   First  of all,  the  status  of  the sampling system is  the same as  it
would be during sampling  of the  flame since these were sequential measurements
and since other gases were sampled  along  with  nitrogen and nitric oxide.   The
other gases presumably were CH/  and 02  although some  CO,  C02,  H2,  and
HoO could be  formed in the hot  tip.   The  presence of  N02 is  probably  due to
Reaction III-3, i.e., NO  + NO +  02  -*• 2NOo.   Using an  analysis similar  to
that described  in  Section III.A.la, this  reaction can convert approximately 5
percent of the NO  to  N02  in the  sampling  system when  sampling the unburned
gases.

     The only remaining possibilities are NO losses  in the flame  zone  or in the
probe.  To shed further  light on this problem,  NO measurements were made at
various levels of  seed for the  three  flames.   These  data are  reported  in Fig.
IV-5.  Examination of these data indicate that  the  loss mechanism is  concentra-
tion dependent.   In the rich  flame,  for example,  63  percent  of  the NO is re-
covered (as NO)  at 1000 ppm whereas  at  2900 ppm only  52 percent  is recovered.
At least in the case  of the rich flame,  it  is  unlikely that  any  NO will  be
converted to  N0? in the sampling line via Reaction  III-3 since virtually no
oxygen exists in the  combustion  products.   These  data are in  fact not  unique
since very similar results have  been  obtained  by  Falcone (1979)  using a  quartz
probe.  For reference, equilibrium concentrations for NO (Gordon and  McBride, 1971)
                                      IV-9

-------
                                                                                FIG. ET-5
                   NITRIC OXIDE MEASURED VS. NITRIC OXIDE SEED
                                         NO. 4 = 0.8
                                         NO. 0 = 1.0
                                         NO. * = 1.2
                                         NOX. ^=0.8
                                         NO CALCULATED ON 'DRY1 BASIS
                                         S.S. TIPPED. WATER-COOLED PROBE
    3000
                                            NOMEASURED = NOSEED
    2000
D_
D.
 Q
 LU
 DC

 C/5
 LU

O
    1000
                                1000
2000
3000
                                         N°SEED -
                                                                               79-10-86-IS
                                          IV-10

-------
at the measured  flame  temperatures  and  for  the  three stoichiometries 0.8,  1.0,
and 1.2 are  approximately  1400,  300,  and  3  ppm,  respectively.   Highest NO
equilibrium  concentrations  are  obtained in  the  lean case  since flame temper-
atures are not that much different  and  since  significantly more oxygen is
present in this  flame  (see  Table IV-A).

     The data presented in  Fig.  IV-5  are  not  sufficient  to identify
which of the two,  i.e., flame  front or  probe  or  possibly  both  are  responsible
for the loss of  NO.  These  data  do, however,  indicate  two facts.   First of all,
the phenomena is concentration  dependent.   (This  fact  is  important  when analyz-
ing data for subsequent optical  measurements  for  which  a  wide  range of NO
concentration was  supplied.)   Secondly,  these results  are not  only  consistent
with possible losses in a water-cooled  probe  (for which  a mechanism does not
exist) but also  are consistent  with a move  towards equilibrium conditions  in
the flame.   Reactions  would only be rapid  in  the  reaction zone of  the  flame
rather than  the  post flame  zone  since concentrations of  reactive species are
the highest  in these regions,  e.g., superequilibrium radical concentrations.
Consequently, losses in the post flame  zone would be minimal which  is  in
agreement with experimental measurements.   Although reaction mechanisms may be
similar to that  suggested  by McCullough,  et.  al.,  (1977),  a detailed analysis
was not considered feasible here due  to the existence  of  very  high  gradients
(temperature and concentration)  and the resultant  difficulty of  incorporating
uncertain transport properties  of reactive  species at  elevated  temperatures.
Although it  could  be argued that in the lean  case,  the equilibrium  concentration
of 1400 ppm  is higher  than  the  lowest concentrations where  losses were  still
observed, it should be remembered that  at  least a  portion  of  the loss  would be
occurring in the early part of  the  reaction zone  where temperatures  and there-
fore equilibrium concentrations  are much  lower.

     Although the  above data are inconclusive, it  is considered most likely
that the loss of NO is primarily due  to flame kinetics rather  than  to  probe
phenomena.   This conclusion has  been  reached  since  the differences  between NO
measured using the different probes is  small  relative  to  the magnitude  of  loss
of NO, especially  and  quite noticeably  in the case of  the  rich  flame.   Similari-
ties between the different  probes are observed in  spite of  noticeable  changes
in surface material and quenching rates.  Furthermore, several  experiments were
performed where  the back pressure was varied  from  about 100 to  600  torr.
Again, these data  reproduced the above  results well  in spite of changes in
quenching rate (see, for example, the discussion  in  Section  III.C.2)  and even
unchoking the sampling orifice.   In addition, experiments  have  been performed
in an adjacent laboratory at UTRC (Seery  and  Zabielski, 1979)  using a molecular
beam/mass spectrometer system  to sample a low pressure (l/10th  atmosphere)
H?/07/Ar flame seeded with  various concentrations  of NO.   In the lean
flames,  NO is quantitatively recovered  in the post  flame  zone,  but  in  the
reaction zone a  profile of  nitric oxide has been  observed,  decreasing  at first
(by as much as 35-40 percent) and then  recovering  back to  the  seed  value.   In
                                     IV-11

-------
the rich cases, initial rates of NO decay are similarly fast but  little  or  no
,ecovery of NO is observed.  These results, which are analogous to  the obser-
vations in this program on a CH./ 02/N2 seeded flame, strongly indicate
that the observed descrepancies are due to flame rather than probe  phenomena.

     Although the data when analyzed in relation to kinetic theory  suggests
that the destruction of NO occurs in the flame and not in the probe,  these
data are not sufficient to draw a conclusion.  The ultraviolet absorption data,
however, which will be reported in TASK III Report are in agreement with the
probe data.  This agreement between data obtained by two separate methods is
sufficient to conclude that destruction does take place in the flame.  Since
the reaction zone where NO loss occurs is extremely small due to  the  high
pressure (1 atm), the present measurements cannot be used to elucidate the
kinetic details of the loss mechanism.  Measurements made at subatmospheric
pressure on the other hand, could prove useful.

IV.A.I.  Uncooled, Stainless Steel Probe

     As described earlier, the uncooled probe is geometrically similar to the
other metallic probes.  This probe, as expected, glowed red when placed in  the
exhaust of the flame.  Using this probe, NO measurements were similar to
measurements obtained when using the cooled probes for the lean flames although
the results were somewhat dependent on the residence time of the uncooled probe
in the flame.  Data are reported in Table IV-B and times between scans are
typically 5-10 minutes.  For the stoichiometric flame, the observed NO was
approximately 25-30 percent less, and for the rich flame values ranged from the
same as that measured using cooled probes to only 20 percent of that value
depending on probe history and probe back pressure,  i.e., residence time.  For
example, at 220 torr back pressure and when the flame is quickly changed from
the lean to rich flame (^ 15-30 seconds), cooled and uncooled probes behave
similarly but in less than a minute the indicated NO begins to fall and after
10 to 15 minutes a stable, but lower value (by a factor 0.66) is obtained.
Then by increasing the back pressure to 430 torr which correspondingly increases
the residence time substantially, the NO drops further to only 20 percent of
the initial value.

     The behavior of the uncooled probe, is not unexpected since  similar results
have been obtained by England, et. al., (1973) and since similar  effects are
common for a stainless steel NOo-NO converter when sampling rich  flame gases.

IV.B.  IFRF Burner

     Initially, temperature profiles were measured and their dependency on
burner operating conditions (i.e., swirl number, position of fuel nozzle, and
design of fuel nozzle) and location within the combustor was examined.  The
primary objectives of these tests were to (1) find stable and repeatable
                                      IV-12

-------
                                 TABLE IV - B
                   MEASURED CONCENTRATION OF NO  (PPM) USING
            UNCOOLED, STAINLESS STEEL PROBE OVER FLAT FLAME BURNER5

height
above
burner
(cm)

1.5
2.0
2.5
3.0
0.81
772
762
755
752
0.82
717
712
710
702
l.O1
655
630
545
500
1.21
242
237
232
230
1.22
227
222
220
215
1.23
225
217
215
215
1.24
147
142
137
132
   Back pressure  225      213      219       218       218      218      435
       (torr)
   Direction of   down     down     down      down      down     up      down
       Scan
   Seed level     971      971      980      1011      1011     1011      1011
      (ppm)

1.  First scan
2.  Second scan
3.  Third scan
4.  Fourth scan
5.  These data may be compared directly to data for  cooled probes presented in
    Figure IV-4.
                                      IV-13

-------
operating conditions and (2) obtain reasonably flat temperature  profiles  in
order to simplify the reduction of the optical data.  The  selected  burner
conditions are described in Section II-C.  Probe locations as  far downstream as
practical were selected to insure that only combustion products  and  not un-
burned or partially burned gases were sampled and that the temperature  profiles
were relatively flat.  Six operating conditions were chosen and  these are
listed in Table II-B.  Two swirl levels were examined and  at each swirl number,
three stoichiometries were tested.  Although flames at lower swirl numbers  were
tested, these flame conditions were relatively unstable and therefore unsuitable
for these experiments.

     A typical temperature profile (corrected for radiation and  conduction)  for
the run condition 41 = 0.8, swirl = 1.25, are shown in Fig. IV-6.  These data
were taken using the aspirated thermocouple described in Section II.C.3.   The
dotted lines represent estimates based on measurements in  the wings.  The
change in swirl produced no measurable difference for any  of the stoichiometries.
The shape of the temperature profiles for the rich and stoichiometric flames
are similar to the lean flame with the centerline temperature varying according
to stoichiometry (see Table IV-C)-  In the wings, temperature measurements
(uncorrected for radiation or conduction) were made using  a chrome1-alumel
thermocouple (0.010" wire diameter) inserted through the open optical ports.
The measured temperatures are much lower than are expected for adiabatic
temperatures of a C-jHg/air flame.  The low temperatures observed are due  to
cooling from the water-cooled walls of the expansion diameter.

     Stable species were measured using the SCOTT Instrument package and  both
the "EPA" and reference probes.  No differences were observed between these
probes and measured concentrations were independent of back pressure.  Experi-
mental data are listed in Table IV-C and equilibrium calculations based upon
the measured (not the adiabatic) temperature are also given.  Data for only
one swirl level is given here since the data for the other swirl numbers  is
essentially identical.  In general, agreement between equilibrium and experi-
mental values are reasonable except for the C02 (and to some extent CO)
values for which the measured values are about 9% low.  It is believed that
this difference is due to uncertainties in the fuel flow rate and/or the  COj
calibration curve.   The high experimental water value (estimated by mass
balance) for the rich flame is due to the presence of about 3.5% molecular
hydrogen (equilibrium value) and the equilibrium value of water  is realistic.
The presence of ^ was not accounted for when estimating the concentration  of
water.

     Concentrations of nitric oxide were measured to be approximately 48, 40
and 25 ppm for the 4> = 0.8, 1.0 and 1.2 flames respectively.  NO  was typi-
cally,  4 to 5% higher than these numbers for the $ = 0.8 and 1.0 flames and  was
not measured for the rich flame.  Since these levels were  too low (even with the
relatively long path length) to provide adequate signal-to-noise ratios in  the
optical measurements, nitric oxide was blended with the inlet air.
                                      IV-14

-------
                                    TEMPERATURE PROFILE ACROSS IFRF COMBUSTOR


                                                       
-------
                                    TABLE  IV-C

                          Mole Percent of  Stable Species
                                 for IFRF Burner1
                                   (Wet Basis)

                                   Experimental
* o22
0.8 4.2
1.0
1.2
CO2
9ppm5
0.063-3
3.0
C022
8.8
10.6
9.0
H203
11.7
15.6
17.3
4
N2 Temp. (K)
75.3
73.8
70.7
1200
1280
1220
                                    Equilibrium

                   02        CO       C02      H20      N2    Temp. (K)
0.8
1.0
1.2
3.92
<1 ppm
•vl ppm
<1 ppm
0.021
3.65
9.45
11.6
9.56
12.6
15.5
13.9
73.1
71.9
68.3
1200
1280
1220
1.  Although two flames were examined (two different swirls) at each
    stoichioraetry, measured values of these stable species were essentially
    the same.

2.  Measured values but corrected for the presence of water vapor.

3.  Water estimated from known input conditions.

4.  Nitrogen calculated by difference.

5.  Error + 40% of value.

6.  Based on measured temperatures

Except where noted, the uncertainity in the experimental concentrations is
approximately ± 5% of the reported (experimental) value.
                                     IV-16

-------
     Typical  profiles  of  nitric  oxide  across  the optical  axis,  normalized
to  the  seed concentration of  NO,  are  shown in Fig.  IV-7.   The profile data
were obtained  using  the  "EPA1  probe and  did not  vary  over the back pressure
range examined (100-350  torr).   The dotted lines are  estimated  extrapolations
based on  other similar flame  measurements.  Data for  the  other  flames are quite
similar.  Also included  on these  profiles  are experimental  data obtained  using
the reference  probe  at two back  pressures.  At these  conditions both the  theo-
retical model  and  experimental pressure  profiles inside  the probe  (presented
in  Section IV.D) indicate that the  flow  in the probe  is  subsonic except  for
possibly  a small region  in the tip.   Thus,  the flow is convectively cooled.
No  differences between NO measurements using  the reference  probe and the  EPA
probe are observed when  operating both in  the convectively  cooled  mode.   For
all presssures examined  and for  both  probes NO  measurements  were  typically
within  2  or 3  percent  of  the  NO measurement.   For Fig. IV-7,  the NO seed  values
for the * = 0.8, 1.0 and  1.2  flames were 184,  189 and 182  ppm dry  and 162,  160,
and 150 ppm wet, respectively.   In  these calculations the  "dry1  concentrations
were estimated assuming  the vapor pressure of water at 3  torr remained in  the
sample  gas with a  total  pressure  of approximately 500 torr.

     Measurements  were also made  when  supersonic flow (verified  by  pressure pro-
files)  extended into the  first constant  area  section of the reference probe.
The data  are  reproduced  in Table  IV-D  and  are compared with measurements made
using the same probe but  at higher back  pressures,  i.e., when the  gases were
convectively  cooled.   Although these  data  are within about  10%,  there appears
to be some difference  between the NO measured at a  low back pressure  (90  torr)
versus  that measured at  higher back pressure.   In addition, the  N0meas/N0seecj
ratios  are smaller than  those obtained at  lower  seed concentrations  and reported
in Fig. IV-7.   This  latter result,  in  fact, is not  surprising since  the results
from the  flat  flame  burner also  show  a concentration dependence.   The former
results apparently indicate a small difference between a  probe  that  convec-
tively  cools  and one that  aerodynamically  cools; however,  it  is  more  likely
that the  observed  differences are associated  with the very  low  operating  pres-
sure of the sampling system.   For example, due to the very  low  pressures,  the
pumps could deliver  only  half (1  cfh)  the  normal flow (2  cfh) to the  CLA.   In
either  case, most  of the  gas  was  extracted with  a 17.5 cfm  vacuum  pump immed-
iately  at the  exit of  the  probe.  Although  the CLA  was recalibrated  to the
lower flow rate, a small  leak of  only  2  to  3% would be difficult to  detect
under normal  flow  conditions  yet  would amount  to a  4-6% dilution when only half
the flow passed through  the sample  line.   In  addition, less water would be
extracted at  the refrigerator since the  total  pressure is  lower.   The resultant
increase in water  concentration will not only act to dilute the  sample on  a
relative basis, but  also  will  provide more efficient quenching  of  the chemilumi-
nescent reaction (see  Section III.A.4) and consequently decrease the  response
of the CLA.   It is estimated  that an  increase in the water  concentration  from
1% to 3% will  decrease the CLA response  (due  to  both chemiluminescence quenching
and sample dilution  effects)  by 4 to 5%.   Consequently, it  is believed that the
differences observed in Table  IV-D are not due to differences in quenching
rates of the gas sample but rather due to  a decrease in sample  line  pressure
and associated phenomena.
                                      IV-17

-------
                             NORMALIZED NITRIC OXIDE PROFILES ACROSS IFRF COMBUSTOR



                                                O 'EPA PROBE


                                                D  1M6IN OO SS TUBE. UNCOOLED



                                                •  REFERENCE PROBE AT 180 TORR


                                                A  REFERENCE PROBE AT 400 TORR
i    o
«    a
    O
        1 4
        1 2
         1 0
        °8
        06
        0.4
        02
        00


-

-40
ffl /->/~>n no o O
j o *o n o o
;
.' ° 2 o o „
t
//
»
1 1 1
-30 -20 -10 C
(j> = 08. SWIRL = 1 25
O ° ^
Q (^=10. SWIRL = 1 25 \
0 o M
00° .'
lj
(^=12. SWIRL = 1 25 II
O 1
o Oo »|
l\l
tf
V
l i l >V
) 10 20 30 4C
                                                    POSITION FROM CENTERLINE (CM)

to
o

-------
                                     TABLE  IV-D

                             Comparison  of  Nitric  Oxide
                       Measurements  Using  the  Reference  Probe
                                     IFRF Burner
NO (ppm)

NOV (ppm)
  A.


NO (ppm)



NO (ppm)
NOX (ppm)
4> = 0.8
Swirl = 1.25
Seed = 960 ppm

   895 ± 40

   880
   830

   820
4> = 1.2
Swirl = 0.63
Seed = 470 ppm

   240 ± 10
                          246
   220
Back Pressure
= 190 torr
(0.25 atm)

Back Pressure
= 380 torr
(0.5 atm)

Back Pressure
= 90 torr
(0.12 atm)1
1.  Supersonic  flow  extends  into  first  constant  area section of  reference
    probe.
                                      IV-19

-------
     Additional information on the performance of  the  reference  probe  is
provided in Section IV.D.I.

IV.C  FT12 Measurements

     Three flight conditions, idle, cruise, and maximum continuous were simu-
lated for this series of tests.  The corresponding operating conditions are
given in Table II-C.  Temperature profiles across  the  optical axis (same  as  for
the IFRF measurements) are shown in Fig. IV-8 for  idle and cruise conditions.
These data were obtained using the Pt/Pt-13% Rh, aspirated thermocouple.   The
profile for maximum continuous is very similar in  shape and magnitude  to  that
for cruise.  Also shown in the figure are data from a vertical profile which
indicate no difference between the different quadrants.  In addition to these
measurements, centerline temperature measurements using coherent anti-stokes
Raman spectroscopy (CARS) were made (Eckbreth, et. al, 1979).  For the CARS
measurements, the temperatures were 580K, 875K and 875K for the  idle, cruise,
and maximum continuous conditions respectively.  These data agree quite well
with the centerline thermocouple data obtained at  the same position i.e.,  590,
900 and 920K.

     Measurements of CO, C02, and 02 using the SCOTT instrument  package
and the EPA probe are listed in Table IV-E.  For comparison, equilibrium data
based on the input conditions and measured gas temperature are also presented.
Good agreement is observed between the experimental and equilibrium data except
for carbon monoxide and an unexplained, high experimental value  for carbon
dioxide at maximum continuous.  The high concentrations of CO measured behind
the FT12 combustors are due to a quenching of the  reaction from  air dilution
within the combustor.  With the reference probe, the COo and 02  concentra-
tions were not measured.  Carbon monoxide concentrations were the same as  with
the EPA probe when the reference probe was operated both in the  convective
cooling mode and with supersonic flow extending into the first constant area
section of the probe.

     Without seed, concentrations of nitric oxide  (total nitrogen oxides)  on
centerline were on the order of 3(15), 5(28), and 6(30) ppm for  the simulated
flight condition idle, cruise, and maximum continuous, respectively.   These
values varied as much as 20-30% from day to day for any given probe but specific
variations due to probe type or back pressure were not observed.  The cause of
the uncertainty may be variations in input conditions for the combustor or
calibration of the CLA at these low NO levels.  Careful attention to obtaining
accurate base line values was not given since this program focused on NO seed
levels much higher than 25 ppm.   In any case, it is interesting  to note that
the N02/NO ratios were typically quite high, on the order of five.  The
source of the N02 is not due to the reaction in the sample line.
                           NO  +  NO  +  02   -*   2 N02
                                      IV-20

-------
                                TEMPERATURE PROFILE DOWNSTREAM OF FT12 COMBUSTOR
                                                         CLOSl.D SYM  IDLE

                                                         OI'IN SYM CRUISE

                                                       O HORIZONTAL PROFILE


                                                       D VERTICAL PROFILE

                                                         (TOP QUADRANT)

                                                      0 A HORIZONTAL PROFILE

                                                         (Cr/AI THERMOCOUPLE)
           1000
               40
30
20
1C
I

O
I
CD
O!
10            0           10

POSITION FROM CENTERLINE (CM)
                                                                20
                                                                                                        30
                                                                            40

                                                                                                                             00

-------
                                     TABLE  IV-E

                          Mole Percent  of  Stable  Species
                                 for  FT12 Combustor
                                   (Wet  Basis)

                                   Experimental
                                   COJ
                                                        H2(r
                 Temp.  (K)
Idle
Cruise
Max. Cont.
0.14
0.19
0.20
19.0
16.8
16.5
0.25
0.20
0.17
1.7
2.8
3.3
1.9
2.6
2.7
77.15
77.60
77.33
580
870
900
                                    Equilibrium
                                  CO       CO-
H20
Temp. (K)
Idle
Cruise
Max. Cont.
0.14
0.19
0.20
17.9
16.8
16.6m
<1 ppb
<1 ppb
<1 ppb
1.94
2.60
2.77
1.94
2.58
2.74
77.3
77.0
77.0
580
870
900
1.  Measured values but corrected for the presence of water vapor.

2.  Water estimated from known input conditions.

3.  Nitrogen calculated by difference.

4.  Based on measured temperatures

The uncertainity of the experimental concentrations is approximately ±  5%
of the reported (experimental) value.
                                      IV-22

-------
since  the  rate  of  this  reaction is  strongly dependent  on the NO concentration
and sample line pressure.   Experiments  with high seed  values of NO ( * 800 ppm)
in air at  much  higher  sample  line  pressure (750 vs  180 torr) indicated only 3%
conversion to N02.   Instead,  it is  more likely that  N02 is formed from NO in the
combustor  during the addition of relatively cold air,  in the post flame region
downstream of the  combustori  or in  the  probe from the  reaction with H0~ (see
discussion in Section  III.A.I).  Insufficient  information is available to
determine  conclusively  which  is the primary mechanism,  however,  it appears
unlikely that probe  reactions are  responsible  due to the relatively low
temperature of  the gas  and  the  necessarily low radical  concentrations.

     Profiles of nitric  oxide were  obtained with seed  levels of  NO at  327,  326,
and 326 ppm for the  idle, cruise,  and maximum  continuous flames.   Normalized
profiles of nitric oxide  for  the idle and  cruise conditions  are  plotted in Fig.
IV-9.  These data were  obtained using the  EPA  probe.   Also shown are profiles
of total nitrogen  oxides.   For  maximum  continuous, the  centerline fraction of
NO (NOX) recovered relative to  the  seed value  was 0.89  (1.08).   In these
gas samples, it is clear  that there are relatively large fractions  of  N02.
For example, idle  conditions  convert more  than 40% of  the total  NO to  N02.  For
the same reasons discussed  in the  previous paragraph,  it is  believed that  the  N02
is probably formed in  the combustor or  post flame zone  rather  than the  probe or
sampling line.   Greater  losses  of NO are observed at idle in spite  of  the  rela-
tively lean stoichiometry possibly  because at  this level mixing  is less intense
and local  variations in  stoichiometry may  be larger  and  may  last  longer than
those at the other power  levels.   In the very  fuel rich  eddies,  losses  similar
to those in the flat frame  burner undoubtedly  take place.

     Also  shown in Fig.  IV-9  are experimental  measurements of  nitric oxide
using the  reference  probe.  For most of these  data the  back  pressure was
approximately 300  torr.  Although good  agreement  is  obtained between measurements
with this  probe and  the  EPA probe,  some differences  are  noted  for one  set  of
NO  measurements at  cruise  and  idle. The FT12  assembly  had been  removed from
the test assembly  and  reinstalled before this  second set of  NO and  NOX
measurements were made.  Small  shifts in alignment of  the fuel nozzle  or
variations  in input  conditions  may  be responsible for  the changes  in NOX
recovery although  no corresponding  change  in the  gas temperature  was observed.

     The reference probe was  also used  to  sample  flame  gases  at  reduced pressure
( «/> 95 torr) where  supersonic  flow conditions extended  into the  first constant
area section of  this probe  (see  discussion in  Section  IV-D).   In  Table  IV-F
these values are compared to  measurements  taken at high  back pressure with  the
same probe.  The agreement  is excellent.
                                      IV-23

-------
            NORMALIZED NITRIC OXIDE PROFILES ACROSS OPTICAL AXIS FOR FT12 COMBUSTOR
o
A,
r
 o
 LLJ
 LU
 
-------
                                     TABLE IV-F
                    Comparison of  Nitric  Oxide  Measurements Using
                         the  Reference Probe -  FT12 Combustor
                                Idle                       Maximum Continuous
Back Pressure           NO       395                               740
 >f 205 torr
(0.27 atm)              N0x      480                               800

Back Pressure*          NO       385                               740
^ 95 torr
(0.13 atm)              NOX      478                               790
* Supersonic  flow  extends  into  1st  constant  area  section  of  reference  probe

IV.D Experimental  Verification  of Probe  Model

     To verify  the predicted  aerodynamic behavior of  gas  sampling  probes,  two
types of measurements  were made.  First, internal pressure measurements were made
for the reference  probe  to  identify if  and when it  operated  in  the aerodynamic
quench mode.  The  smaller  probes were not  instrumented  due to physical limitation.
Secondly., mass  flow measurements were made for probes of  three  sizes, macro-,
mini- and microprobes, to  compare with  theoretical  values.   These  data include
not only measurements  of the  maximum possible  flow  for  a  given  probe but also
the variation of mass  flow as a function of  probe back  pressure.

IV. D.I  Pressure  Profiles  for  the  Reference Probe

     To verify  the operation  of the reference  probe,  three static  pressure taps
were positioned along  the  constant  area  section and one was  placed after the
bend.  A detailed  design of the probe including the location of  these  pressure
taps is given in Fig.  III-l.  The last  pressure tap was used to  approximate the
stagnation pressure at the exit of  the  probe and  this piece  of  data is of  prime
importance in estimating the  location of shock recovery.  Typical  static pres-
sure data obtained  when  sampling the exhaust from the FT12 combustor are shown
in Fig. IV-10.  The solid  and dotted lines in  the figure  are calculated profiles
when the stagnation pressure  at the exit of  the probe (back  pressure)  was
assumed to be equal to the static pressure after  the  bend.   For  these  calcula-
tions,  this assumption can be shown to  provide a  reasonable  estimation.  As can
be seen, the agreement between  the  theoretical and  calculated profiles is
excellent.  The mechanism  for reducing  the static pressure to the  measured
value was a normal  shock whose  position  and, therefore, strength varied with
back pressure.
                                      IV-25

-------
                                 PROFILES OF STATIC PRESSURE FOR THE REFERENCE PROBE
                              TEXTERNAL = 90° K




                              PEXTERNAL" 1 8tm
SOLID LINES — CALCULATED PROFILES



SYMBOLS — EXPERIMENTAL MEASUREMENTS
I.U
0.8
s
i-
3 ^ 0.6
1 UJ
KJ (E
_,

—


* i
<£ 1 BACK PRESSURE 0 46 aim
£ 1 >x* — "" "^
o 0.4 [- /^
< 1
^

.2

0
1 022 aim ^ ^
\ °
\ 013 aim
N - ,,^ 	 CHOKED ^
X_ , n 1 	 J 	 1 	 IT) 	 * ' 1 1 " 1















0 2 4 6 8 10 12 14 16 18 20 22
POSITION FROM TIP (cm) Z!
0
M

-------
     The  calculations  indicate  that  at  the lowest back pressure the flow
 chokes  prior  to  the  exit  of the constant area section.  Since the flow in a
 constant  area tube can choke  only at  the exit of the tube,  the predicted choke
 location  is  in error due  apparently  to  uncertainties in the back pressure and
 the  skin  friction correlation used in the analysis.   In any case, the experimen-
 tal  data  presented  in  Fig.  IV-10 confirm the analytical result that an extremely
 low  back  pressure is needed to  extend supersonic flow for a substantial distance
 into the  probe.

     Experimental verification  of calculated temperature profiles was considered
 impractical since the  presence  of a  thermocouple in  the supersonic  region will
 trigger a  shock  wave.   Even if  placed downstream of  the recovery shock,  thermo-
 couple  temperatures  would probably be ambiguous  since the thermocouple  will
 also be cooled by its  water-cooled jacket.   Nevertheless,  some temperature
 measurements  were made using  the chromel-alumel  thermocouple  depicted  in  Fig.
 III-2.  Although the measurements typically  were within 50K of the  theoretical
 calculations,  this agreement  may be  fortuitous.

     NO measurements taken  with the  reference probe  operating at  low  back pres-
 sure are  given earlier in this  chapter.   Typically,  for these measurements, the
 back pressure was 0.12 atmospheres.   Under these conditions,  the  very  low
 pressure measurements  at  the  first pressure  tap  verified that  supersonic  flow
 exists  at  least  past this position.   For the IFRF burner at  4> =  1.0,  a  computer
 analysis  indicates that the shock occurred 5.2 centimeters  from  the probe tip
 (45 microseconds).   The static  gas temperature recovered to  1080K (compared to
 the  external  flame temperature  of 1280K) and cooled  below 1000K  in  a  total time
 of 170 microseconds.  Alternatively,  at  a back pressure of  1/3 atmosphere, the
 computer  calculations  indicate  that  the  shock occurs  0.9 cm from the  tip  (8.6
 microseconds).   After  this  shock,  the static temperature increased  to  1205K but
 did  not  fall below  1000 K  until 750  microseconds.   Clearly;  there  is  substan-
 tial difference  between these two modes  of operation.   When the  gas flow  was
 supersonic past  the  first pressure tap,  the  time above  1000K  was  nearly 6 times
 less than  that when  the flow  was primarily subsonic  (0.125 vs  0.74 milliseconds).

     It was desirable  to extend  the  aerodynamic  quench  made by reducing the back
 pressure  further  and simultaneously measuring the nitric oxide concentration;
however, this  was not  possible  because  of the low pumping speed  of  sampling
 pumps operated at low  pressures.   At  a back  pressure  of 90  torr  (0.12  atm), the
maximum flow  rate through the sampling  lines was only one cubic  foot/hour
 ( s 7.9 sec/sec)  which  was barely sufficient  to operate  the  chemilurainescence
 analyzer.  Only  this amount of  flow could be obtained  in spite of the  use of
many metal bellows pumps  (see Section II.D.la) in the gas sample  line  and the
vacuum pump associated with the  CLA.  To extend  the  supersonic flow all  the way
 through the constant area section, the  pressure  taps  indicated that the back
pressure had  to  be reduced  to ^ 50 torr.   At this low sampling line pressure,
 the collection of sampling  pumps would  not supply sufficient  flow to  the  CLA.
                                       IV-2 7

-------
IV. D.2  Mass Flow Measurements

     Measurements of mass  flow through  the micro-, mini-,  and macroprobes  were
made using low pressure drop, Hastings mass  flow meters  (ALU-100,  -5K,  and -20K,
respectively).  Description of the experimental method used  for  these measure-
ments is provided in Section II-E.  Hot  flow data were obtained  for  each class
of probes and cold flow measurements were made only  for micro- and miniprobes.
For the macroprobe at room temperature,  the mass flow was  too large  for accurate
readings on the Hasting meter (mass flow rate through probe  
-------
                                                                              FIG. IV-11
      RELATIVE MASS FLOW VS. BACK PRESSURE FOR SEVERAL PROBES
                MICROPROBE
               (75
-------
mass flow profiles exhibit some differences between the cold and hot  flows;
however, these differences cannot be explained using our current understanding
of the fluid mechanics and heat transfer within probe tips.  Additional research
is required to identify the controlling phenomena.

IV.D.2c  Microprobe

     Measurements using a quartz, water-cooled microprobe with an orifice diam-
eter of 75 microns (0.003 inches) were also made over the flat flame burner
( $ = 0.8) and for room temperature air.  A photograph of a similar probe is
shown in the Task I report (Dodge, et. al, 1979).  Normalized mass flow profiles
are provided in Fig. IV-11.  For the cold flow, the data indicate that the flow
chokes for all back pressures below (approximately) one-half atmospheres.  This
choking point is consistent with not only the experimental data from the macro-
and miniprobes, but also with equations  developed by Shapiro (1953) for the
required pressure drop for choking at a minimum area.  Air, for example, will
choke with a pressure ratio (static pressure at minimum area to stagnation
pressure) of 0.53.  This ratio is relatively independent of changes in gas
properties.  For the hot flow data, a substantial difference is observed
when compared to the other results.  The flow does not choke until at least
a back pressure below one-fifth of an atmosphere.  This result is in qualitative
agreement with the computer calculations (for probe design) that indicated that
this microprobe cannot be choked (see Section III.C.2a).  Based on both an
analytical study and experimental measurements, one must conclude that under
described conditions the flow does not choke (except possibly at very low back
pressure) and  consequently will not quench the gas aerodynamically.  This
conclusion is in direct conflict with assumptions regarding probe behavior in
other investigations and suggests that some earlier conclusions based on
(assumed) very high rates of quenching may have to be reanalyzed.

     The experimental and analytical results for the microprobe do not agree
as well as for the larger probes.  In addition to the uncertainties mentioned
previously, it should be noted that the dump loss calculation procedure used
here becomes singular at a high area ratio.  The curve in Fig. IV-11 was
calculated for the highest area ratio prior to the breakdown of the calculation
procedure (^550), while the actual area ratio was approximately 4500.  The
results for the cold flow condition indicate that a choked orifice may exist
over a range of back pressure if the appropriate sudden expansion loss is
applied.  On the other hand, the sudden expansion model for this loss is not
adequate to explain the variation in the hot flow data which indicates that the
probe orifice is not choked.

IV.D.  3  Discussion
     The experimental mass flow vs. back pressure profiles indicate that, except
for the microprobe operated at high gas temperatures the probe orifices become choked
                                      IV-30

-------
at the proper back pressure.  The back  pressure  to  ambient  pressure  ratio  at
which choking occurs is approximately 0.5  and  is  consistent with  calculations
for choked flow through an orifice plate (Shapiro,  1955).   The  probe  design
model assumes that the flow within the  probe  is  not  separated.  For  the  probes
examined in this  study, however, separated  flow  occurs when the probe is
operated unchoked.  Except in the case  of  the  microprobe  operated  in  a high
temperature gas stream, calculations accounting  for  sudden  expansion  losses
yield mass flows  in agreement with the  measured  data.

     It should be noted that the probe  design  model  is generally  used to design
probes for which  a sudden expansion  loss is not  important.  Either the flow is
supersonic throughout the probe tip  (as in  designs  exploiting an  aerodynamic
quench) or the tip area ratio and flow  angle  are  small enough that the subsonic
flow in the tip does not separate.   Microprobes  are  designed to minimize the
disturbance that  the probe causes to the flame front.  Unfortunately,  the
geometry of a microprobe requires a  better  understanding  of its internal flow
characteristics than heretofore appreciated or included in  the  analysis.
                                   IV-31

-------
                           V.   RESULTS AND DISCUSSIONS
      In  order to study the processes involved in the probe sampling of NO
 and,  ultimately, to compare probe and optical measurement methods properly,
 three combustion devices  were designed and characterized under a variety of
 operating  conditions.   The most  controlled of these devices was a flat flame
 burner.  Temperature and  NO concentration profiles for methane/nitrogen/oxygen
 flames were  obtained at various  heights above this burner.  The stoichiometries
 characterized were  0.8, 1.0,  and 1.2.  Centerline temperatures in the  post-flame
 regions  varied from 1740  K to 1815 K depending on stoichiometry and height.
 Thermal  NO produced in these  flames ranged from 5 ppm to 30 ppm.   Since these
 concentrations were considered too low for making accurate optical  absorption
 measurements,  NO was seeded into the flame at concentrations of 850 ppm to 4500
 ppm.   The  amount of seeded NO recovered in the post flame region  was dependent
 on  stoichiometry and will be  discussed below.

      The second device employed  was a swirl burner designed after that of the
 International Flame Research  Foundation (IFRF).   Temperature and  NO concentration
 profiles were obtained for unseeded and NO-seeded propane-air flames.   The
 stoichiometries used were 0.8, 1.0, and 1.2.   In addition, tangential  or swirl
 flows  were introduced  onto the air flow.   Swirls (S)  of 0.63 and  1.25  were used
 where  swirl  is defined as the ratio of tangential momentum to axial  momentum
 divided  by the radius  of  the  exit.   Peak  temperatures  measured  by suction
 pyrometry  and corrected for radiation and convection  ranged from  1200  to 1280 K.
 Thermal  NO concentrations ranged from 25  to 48 ppm and were principally dependent
 on  stoichiometry.   Seeded concentrations  were varied  from 130 ppm to 890 ppm.
 The amount of NO recovered was also dependent on stoichiometry.   Temperature
 and concentration profiles were  obtained  87.5 cm downstream of  the  quarl.  Data
 obtained in  four quadrants indicated axial symmetry.   The bulk  of exhaust was
 contained  in  an expansion chamber of 50.0 cm diameter.  To simplify  optical
 access,  windowless  ports  with purging across the exit  of the optical ports were
 used. Temperature and  concentration profiles  through  and external to the ports
 were  obt ained .

      The third combustor,  which  was also  installed into the above expansion
 chamber, was  a modified Pratt &  Whitney FT12  combustor.   The modifications
 consisted  of  a reduction  in length from 41.1  cm  to 29.5 cm and  the  closing of
 several  air holes to provide  axial  symmetry.   The distance from the  combustor
exit  to  the probe (optical  axis)  was 78.0 cm.   The combustor was  operated at
 three conditions:   idle,  cruise,  and maximum  continuous.   The fuel/air ratios
were 0.0106,   0.0143, 0.0152,  respectively.   Operating  air pressure  was Mach
number scaled  from  6.0 atm to 1.0 atm.  The scaling was performed because of
 facility limitations and  to facilitate  accurate  definition of the optical path.
Centerline temperatures for the  three operating  conditions were 580  K, 870 K,  900  K,
respectively.   Thermal NO produced  spanning the  range  from 3 ppm  to  6  ppm, while
                                       V-l

-------
total N0x values ranged  from  12  to  35 ppm.  In  order  to  obtain  adequate
signal-to-noise ratios  in  the optical measurements,  the air  flow was  seeded
with NO  in concentrations  from 270  ppm  to 790  ppm.   As  with  the  flat  flame
burner and IFRF combustor, the amount of NO recovered was dependent on fuel/air
ratio ( stoichioraetry).

     The measurement  of  NO by extractive sampling  is a  procedure consisting of
three elements each  of which must be considered  in detail.   The  first  of these
is the removal of the sample  from the flame without  seriously  perturbing the
flame while rapidly  reducing  sample temperature  and  pressure.  The second  element
involves the transfer of the gas from the probe  to the  analyzer.  This  transfer
includes, usually, the removal of water and particulates.  The third element  is
the actual analysis  of NO  with a chemiluminescent analyzer.  This study  has
been concentrated on  the first two  elements.   However,  information obtained
outside  this particular  study has been  applied to the third  element.

     Two separate sets of  probes were designed and evaluated and  set for the  flat
flame burner measurements  and the other set for  the combustor  tests.   For  each
set of probes, a special effort was made (using  a computer code  for probe
design)  to select a  probe  geometry  that was capable of  cooling the gas aero-
dynamically.  For the flat flame burner, the probes used were:   (1) a water-
cooled quartz miniprobe; (2) a water-cooled, stainless  steel miniprobe;  (3)  a
water-cooled, stainless  steel miniprobe with a copper tip; and (4) an uncooled
stainless steel miniprobe.  Each of these probes had an  orifice  diameter of  635
V , and the metallic  probes had an area  ratio of  16.  The area  ratio for  the
quartz probe was 62.  In addition,  a water-cooled quartz probe with an orifice
of 75 V  and an area  ratio  of 4500 was also  examined, however, only mass  flow
measurements were made with this probe.  For the large  combustor, two water-
cooled stainless steel macroprobes  were designed.  The  first of  these, which
was defined as the EPA probe  (in compliance with the Federal Register  require-
ments), had an orifice diameter equal to 0.080 in (2 mm) and an  area ratio  of
2.4.  The second macroprobe, which  was  defined as the reference  probe, had  an
orifice diameter of  0.080  in  (2 mm) but with dual expansion  areas.  The  area
ratio of the first relative to the  tip  was  10.6  while the second  ratio relative
to the t ip was 29.7.

     In addition to properly sizing the probe  to minimize temperature  and  flame
perturbation,  these probe  designs were  analyzed  in relation  to the available
kinetics of NO decomposition and the aerodynamics necessary  to achieve quenching
of NO react ions.

     A consideration of gas phase reactions througi,  'hich NO could be destroyed
revealed that  if the gas sample was rapidly cooled ( •/* 1 msec)  to  below 1000  K,
NO loss would not occur.   In addition,  it was  estimated  that catalytic effects
would be minimized if exposure to walls whose  temperature was greater  than 600  K
was kept at 10 usec or less.  This  kinetic  analysis  is  consistent with the
results of this study obtained with probes  satisfying the above  conditions.  The
                                       V-2

-------
aerodynamic behavior of these probes were  studied with a computer model based
on the original work of Cohen and Guile  (1970).  This program predicts the
changes in static temperature and pressure  as  a  function of  time and  position
in the probe.  Parameters considered were  area changes, heat transfer rates,
skin friction, and probe back pressures.   In addition, mass  flow is calculated
for specified combustion stream temperatures and pressures,  geometries, and
back pressures.  The most important prediction of this model was that micro-
probes and miniprobes  (as defined above) cannot  achieve an aerodynamic quench.
In such a quench, the  gas is supersonically expanded to reduce  the static
temperature and pressure rapidly but does  not  shock heat until  the stagnation
temperature is significantly reduced below  source conditions via convective
cooling to the walls.  Cooling sufficiently rapid relative to kinetic rates,
however, can  still be  achieved only by convective cooling to the walls under
subsonic flow conditions.

     For the  macroprobes (large scale),  a  study  of various geometries for
sampling from many temperatures, revealed  that an aerodynamic quench  is pos-
sible.  The reference  probe was constructed from the model predictions.  The
design for the EPA probe was selected  for  rapid  cooling by convection.  The
significant results obtained will be reviewed  below.

     The second element of NO extractive sampling, i.e., sample transfer, can
be subject to two related problems.  The first is the conversion of NO to
N02.  The second is the loss of NOo once converted from NO.  The principal
mechanism for the conversion is

                             NO + NO + 02 -» 2N02

This conversion is related directly to the  partial pressure of 02 and to
the square of the partial pressure of NO present.  Since the reaction is
relatively slow, conversion can be minimized by minimizing the sample transfer
time.  For laboratory  flames, this reaction is only important for lean flames
and, more critically,  for high seed levels  of NO.  For large scale combustors
which typically have an overall lean stoichiometry. the above reaction is
important.  Federal Register (1976) regulations  for sample transfer time (2
seconds) seem adequate.  A vital part of the sample transfer process, however^
is the time spent in the analytical instrumentation.  For some instruments,
this time can be considerable.  An example  of  this problem was considered and
analyzed experimentally.  The rate coefficient for the above reaction was
obtained with good agreement ( ^ 50%).  The  Federal Register (1976) requires 9
seconds to 15% response for N0x-  For low values of NO (-M00-200 ppm) this
seems adequate; however, for high concentrations (> 1000 ppm) careful considera-
tion of this  reaction  is important to the measurement of NO.   If conversion of
NO to N02 occurs, then NO loss (NOX loss)  is possible in water traps and
particulate filters as has been previously  documented (Tuttle, et.  al., 1973).
                                       V-3

-------
     The final element to be addressed is sample analysis with a chemiluminescent
detector.  For laboratory flames with unusual carriers such as He and Ar,  sig-
nificant errors in calibration are possible due to quenching differences  and
viscosity effects associated with sampling handling.  These problems are
avoided by having calibration gases gravimetrically prepared in the desired
diluent and, if possible, analyzed by an alternative method.  For laboratory
and cotnbustor flames in air, measurements made on wet samples can have errors
in excess of 10%.

     The probe measurements of seeded-NO over the flat-flame burner indicated
the following.  First, within the random errors of the measurements, no signi-
ficant differences were observed between the water-cooled quartz, water-cooled
all stainless steel, and water-cooled copper tipped stainless steel probes even
though there was significant visible radiation from the all stainless steel
probe and none from the copper tip.  A range of back pressures (100 torr  to 700
torr) were employed but the results were independent of this parameter.   The
uncooled stainless steel probe produced results similar to the water-cooled
probes in the lean flame.  For the stoichiometric and rich flames, NO results
were significantly lower.  In addition, for all stoichiometries a systematic
difference between the NO-seed value and the NO measured by the water-cooled
probes was observed.  These differences were the greatest for the rich flame
while for the lean and stoichiometric flames the differences were approximately
the same.  For all stoichiometries, the amount of NO destruction was dependent
on seed concentration.  This fact by itself was not sufficient to conclude that
the NO was being destroyed in the flame and not in the probe.  However, the
optical measurements obtained in TASK III of this study indicate that the  loss
is occurring in the flame front.  The specific reaction for this loss was not
analyzed.  It is possible, nevertheless, to suggest that the NO concentration
is slowly driving towards equilibrium.  Similar results were observed on  the
IFRF diffusion flame.  Here, as with flat flame, the amount of NO- observed
cannot explain the loss.  For the FT12, a loss was observed forx all operating
conditions.  Although these conditions are lean overall the locally rich
regions in the combustor are responsible for the loss.

     The macroprobes used in making the NO measurements on the large scale
combustors revealed no significant differences in detected NO.  The reference
probe was operated both in the aerodynamic quench mode and in the convective
quench mode with no significant difference in detected NO.  It is important to
note that the low back pressure required to operate a probe in the aeroquench
mode is difficult to achieve.  The results do not warrant the effort necessary
to make these measurements.

     Finally, the validity of the computer model for the probe design model was
investigated.  Pressure measurements made internal to the reference probe
verified the predictions of this model and demonstrated that a probe could be
operated in an aerodynamic quench mode albeit at extremely low back pressures.
                                       V-4

-------
Mass flow measurements  for varying back pressures were  in  agreement with  the
predictions  for the mini- and macroprobes when  sudden expansion  losses  were
considered.  Mass  flow measurements on the microprobe indicated  that when
sampling at  atmospheric pressure and high temperatures  (^ 1800 K)  the  flow in
the probe is not choked until at least below  0.2 atm.   This  is significantly
lower than that predicted by cursory analysis  (^ 0.5 atm).   The  nonchoking of
the probe was accurately predicted with the detailed fluid flow  analysis
contained in the probe design model.
                                       V-5

-------
                           VI.  CONCLUSIONS
     Based on the results of this study,  the  following major  conclusions  can  be
drawn for well designed and properly operated  probes  sampling the  exhaust  gases  of
gaseous and liquid fueled combustors.

     1.  Water-cooled quartz, stainless  steel, and  copper-tipped miniprobes
         yield the same NO concentrations when sampling products from methane/
         oxygen/nitrogen flames with seeded NO at temperatures up  to 2000  K and
         irrespective of stoichiometry.   This  similarity  in behavior occurs
         over a wide range of probe back  pressures; hence, no advantage is
         gained from back pressures less  than  0.5 atm when sampling atmospheric
         pressure flames.

     2.  Uncooled stainless steel probes  give  NO concentrations slightly  smaller
         (10-15%) than the cooled probes mentioned  above  for  lean methane/oxygen/
         nitrogen flames (1>1800K).  For  stoichiometrie and rich flames,  the  NO
         concentrations are significantly less and, at least  for the rich
         flames, the amount of loss is dependent on the probe  back pressure.
         The destruction of NO in this uncooled probe is  similar to that
         encountered in NO/NOj converters operated  in the absence of oxygen.
         Hence, uncooled stainless  steel  probes are only  suitable  for sampling
         NO in the presence of oxygen.

     3.  In general, for miniprobes and microprobes, aerodynamic quenching is
         not possible because of fluid mechanical and geometric constraints.
         However, rapid-cooling of  the sample  gases (within a  few milliseconds)
         can be achieved by convective heat transfer to the probe walls.

     4.  For microprobes, mass flow measurements indicate that, in the sampling
         of high temperature gases, choked flow does not  occur at the classical
         pressure ratio.  This result was predicted by the probe model and
         suggests that quenching processes may be less efficient than those
         estimated in previous studies.  A kinetic analysis indicates that
         quenching rates are still  sufficiently fast for  sampling NO in exhaust
         gases; the impact of slower quenching rates on gases  sampled from
         reactive flame zones may have to be examined.

     5.  Unlike the small scale probes,  it is possible to construct a large
         scale water-cooled probe that can produce aerodynamic quenching of the
         sample; however, measurements made on both gaseous and liquid fueled
         combustion systems yielded essentially the same  results regardless of
         the quenching mode, i.e., aerodynamic or convective.  Given this  fact
         plus the complexity of probe construction and the difficulties in
         achieving the low probe back pressures required  of the aerodynamic
         mode,  there is no advantage to aerodynamic quenching  in the measure-
         ment of NO.
                                       VI-1

-------
The model predictions of pressure distribution and mass flow within
the large scale probe agreed well with the experimental data.  In
addition, the model accurately predicted that the microprobe would not
choke at the classical pressure ratio when sampling at high tempera-
tures.  Based on these results, it can be stated that the fluid dy-
namic and heat transfer processes have been adequately described in
the model for the case of fluid mechanical choking at the probe
ori fice.

A kinetics analysis of gas phase reactions known to destroy NO indi-
cated that no significant loss of NO would occur during the sampling
process if the sample temperature was reduced to 1000 K in approxi-
mately 1-2 milliseconds.  The results of this study are consistent
with this analysis.

A review of the literature indicated no definitive study where large
differences (> 20%) between measurements of total nitrogen oxides from
different probes were observed when properly designed probes and
sampling lines were used and correct calibration of the chemilumi-
nescent analyzer was performed.  A properly designed probe is one that
does not perturb the flame environment, does not stagnate the flow, is
water-cooled, operates at a back pressure to external static pressure
ratio low enough to aid in quenching of reactions, and finally has hot
walls (> 600 K) which are limited only to the front portion of the tip
such that the local gas residence time is on the order of 10 micro-
seconds or less.  Proper sampling lines are those in which the resi-
dence times are short relative to the time required for the conversion
of NO to N0« and the loss of NO  in water traps and particulate
filters.  Correct calibration of the chemiluminescent analyzer con-
sists of accounting for the influence of gases other than N? on the
introduction of NO into the chemiluminescent reaction chamber and the
collisional deactivation of excited N02-
                             VI-2

-------
                                  REFERENCES


Allen, J. D.:  Combustion and Flame, 24, 133 (1975).

Amin, H.:  Combustion Science and Technology, 15, 31 (1977).

Baulch, D. L. , D. D. Drysdale, and D. G. Home:  "Critical Evaluation of
Rate Data for Homogeneous, Gas Phase Reactions of Interest in High Temperature
Systems".  The University, Leeds, England, 1970.

Baulch, D. L. , D. D. Drysdale, D. G. Home, and A. C. Lloyd:  Evaluated Kinetic
Data for High Temperature Reactions, Vol 2, Butterworth, London, 1973.
  i                                                                      i
Beer, J. M. and N. A. Chigier:  Combustion Aerodynamics, Ed. by J. M. Beer,
John Wiley and Sons, Inc., New York, 1972.

Benson R. and G.  S.  Samuelsen:  Presentations at the Western State Section of
the Combustion Institute, Fall Meeting, Paper No. 76-39, October 18-20, 1976
and Spring Meeting,  Paper No. 77-7, April  18-19, 1977.

Benson,  R., G. S. Samuelsen and R. E. Peck:  Presentation at the Spring
Meeting Western State Section of  the Combustion Institute, Paper No. 76-11,
April  19-20,  1976.

Beal, J. L. and Grey, J. T.:  Journal of the American Rocket Society 23,
p. 174 (1953).

Bilger,  R. W.:  Probe Measurements in Turbulent Combustion.  PURDU-CL-75-02,
Purdue University,  School of Mechanical Engineering, May 1975.

Bilger,  R. W. and R. E. Beck:  Fifteenth Symposium (International) on
Combustion, The Combustion  Institute, Pittsburgh, 541 (1975).

Bryson,  R. J. and J. D. Few:  "Comparisons of Turbine Engine Combustor
Exhaust  Emissions Measurements Using Three Gas-Sampling Probe Designs",
AEDC-TR-78-7, November  (1978).

Cernansky, N. P.:   AIAA Aerospace Sciences Meeting, Paper No. 76-139,
Washington, D. C.,  January, 1976.

Cernansky, N. P.  and S. Singh:  "Sampling  and Measuring NO and N02 in
Combustion Systems".  Final Report for NSF Grant No. 76-22112, Drexel
University, April 1979.

Chen, J. Y.,  W. J.  McLean and F.  C. Gouldin:  Paper No. 79-17 presented
at the Western States Section of  the Combustion Institute, Spring Meeting,
April  1979.
                                       R-l

-------
                              REFERENCES (Cont'd)
CIAP (Climatic Impact Assessment Program) DOT-TST-75-51, 52 (1975).

Clark, J. A. and A. M. Mellor, Paper presented at the Gas Turbine
Conference and Products Show, ASME, Paper No. 80-GT-71, March, 1980.

Cohen, L. S. and R. N. Guile:  AIAA Journal, 8^, 1053 (1970).

COMESA (Committee on Meterological Effects of Stratospheric Aircraft)
U.K. Meterological Office, Bracknell (1975).

Cornelius, W. and W. R. Wade:  Transactions of Society of Automotive
Engineering, Paper No. 700708, 1970.

COVOS (Comite d'Etudes sur les Consequences des Vols Stratospheriques)
Societe Meterologique de France, Boulogne (1976).

Crutzen, P. J.:  Quart. J. Royal Met. Soc., 96, p. 320 (1970).

Crutzen, P. J.:  Ambio, 1_, p. 41 (1972).

Davis, M. G., W. K. McGregor and J. D.  Few:  Arnold Engineering Development
Center Report AEDC-TR-74-124 (AD-A004105), (1976a).

Davis, M. G., J. D. Few and W. K. McGregor:  Arnold Engineering Development
Center Report AEDC-TR-76-12 (ADA021061), (1976b).

Davis, M. G., W. K. McGregor and J. D.  Few:  J. Quant. Spectrosc. Radiat.
Transfer, 16, p. 1109 (1976).

Dimitriades, B.:  Journal of Air Pollution Control Association, 17, p. 238
(1967).

Dodge, L. G., M. B. Colket, M. F. Zabielski, J. Dusek and D. J. Seery: "Nitric
Oxide &  Measurement Study:  Optical Calibration".  Report prepared  for DOT/FAA
Contract No. FA77WA-4081 (1979).

Dodge, L. G., M. F. Zabielski, L. M. Chiappetta and D. J. Seery: United
Technologies Research Center, Unpublished report (1979a).

Eckbreth, A. C., P. A. Bonczyk and J. F. Verdieck:  "Investigations of CARS and
Laser-Induced Saturated Flourescence for Practical Combustion Diagnosis".
Final ki-port under EPn contract 68-02-3105, United Technologies Research Center
K79-954403-13, September, 1979.

LcKtrt,  L. K. b. and R. M. Drake:  Heat and Mass Transfer (New York; McGraw-Hill,
1^5^), pp. 15--158.

                                        R-2

-------
                               REFERENCES  (Cont'd)
England,  C.,  J.  Houseman  and  D.  P. Teixeira:   Combustion  and Flame
2,0,  p. 439,  (1973).

Falcone,  P. K.,  R. K. Hanson,  and C.  H. Kruger:   Paper No. 79-53 presented
at the Western States Section of the  Combustion  Institute, October 1979.

Falcone,  P.:  Mechanical  Engineering  Department,  Stanford University, Private
Communication, 1979.

Federal Register, Vol.  38,  No.  136, July  17,  1973, Vol. 41, No. 181,
September 16, 1976.

Few, J. D., R. J. Bryson, W.  K.  McGregor  and  M.  G. Davis:  in Proc. Intl.
Conf. Environmental  Sensing and  Assessment, Las  Vegas, Nev. (Sept. 1975).

Few, J. D., R. J. Bryson, and  W. K. McGregor:  Arnold Engineering Development
Center Report AEDC-TR-76-180  (1977).

Flower; W., R. K. Hanson, C.  H.  Kruger:   Fifteenth Symposium (Int'l) on
Combustion.   The Combustion Institute, p. 823 (1975).

Folsom, B. A. and C. W. Courtney:  Proceedings of the Third Stationary
Source Combustion Symposium Volume II Advanced Processes and Special Topics.
EPA  60017-79-050b, Feb.,  1979.

Friedman,  R.  and J.  A.  Cyphers:  Journal  of Chemical Physics,  23,  p.  1875,
(1955).

Fristrom,  R. M.  and  A.  A. Westenberg:  Flame  Structure, McGraw-Hill,  New
York, 1965.

Glawe, G. E., F. S.  Simmons,  and T. M. Stickney:   NACA TN 3766, 1956.

Gordon, S. and B. J. McBride:  "Computer  Program for Calculation of Complex
Chemical  Equilibrium Compositions, Rocket Performance, Incident and Reflected
Shocks, and Chapman-Jouquet Detonations".  NASA  SP-273 (1971).

Gryvnak,  D. A., D. E. Burch:   AFAPL-TR-75-101, (1976a).

Gryvnak,  D. A., D. E. Burch:   AIAA Fourteenth Aerospace Meeting, AIAA No.
76-110, Washington, D.  C.,  (I976b).

Halpern,  C. and F. W. Ruegg:   Journal of  Research of the National  Bureau
of Standards, 60, p. 29 (1958).
                                      R-3

-------
                              REFERENCES (Cont'd)
Halstead, C. J., G. H. Nation, and L. Turner:  Analyst 9_7, p. 55, (1972),

Hanson, R. K., W. F. Flower and C. H. Kruger:  Combustion Science and
Technology, £, p. 79, (1974).

Milliard, J. C. and R. W. Wheeler:  Combustion and Flame 29^, p. 15.
(1977).

Jensen, D. E. and G. A. Jones:  Combustion and Flame 32, p.l, (1978).

Johnson, G. M. , M. Y. Smith and M. F. R. Mulcahy:  Seventeenth Symposium
(International) on Combustion, The Combustion Institute, Pittsburgh,
p. 647, 1979.

Johnston, H. S.:  Science, 173, p. 517.

Kays, W. M.:  Numerical Solutions for Laminar-Flow Heat Transfer in
Circular tubes, Transactions ASME, p. 1265, November 1955.

Koshi, M. and T. Asaba:  International Journal of Chemical Kinetics,
9^, p. 305. (1979).

Kramlich, J. C. and P. C. Malte:  Combustion Science and Technology,
J^, p. 91, (1978).

Land, T. and R. Barber:  Transactions of the Society of Instrument
Technology £, p. 112, (1954).

Lengelle G. and C. Verdier:  Gas Sampling and Analysis in Combustion
Phenomena.  AGARD-AG-168, July, 1973.

Lyon, T. F., W. C. Colley, M. J. Kenworthy and D. W. Bahr:  Development
of Emissions, Measurement Techniques for Afterburning Turbine Engines.
AFAPL-TR-75-52 , October, 1975.

Matthews, R. D., R. F. Sawyer, and R. W. Shefer:   Environmental Science,
and Technology, 11, p. 1092, (1977).

McCullough, R. W., R. H. Kruger, and R. K. Hanson:  Combustion Science
and Technology J_5_, p. 213, (1977).

McGregor, W. K., J. D. Few and C. D. Litton:  Arnold Engineering
Development Center Report AEDC-TR-73-182, AD-771  642 (1973).
                                      R-4

-------
                               REFERENCES (Cont'd)


 McGregor,  W.  K., B.  L.  Seiber and J. D. Few:  in Proc.  Second Conf.
 Climatic  Impact  Assessment Program, Cambridge, Mass. (Nov., 1972).

 Meinel, H.  and L.  Krauss:   Combustion and Flame _33, P-  69 (1978).

 NAS  (National  Academy of Science), "Environmental Impact of Stratospheric
 Flight",  Climatic  Impact Committee (1975).

 Oliver, R.  C., E.  Baver and Wasylkiwskyj:  DOT-FAA-AEE-78-24.

 Oliver, R.  C., E.  Baver, H. Hidalgo, K. A. Gardner and  Wasylkiwskyj:
 DOT-FAA Rpt.  FAA-EQ-77-3 (1977).

 Rohsenow,  W.  M.  and H. Y.  Choi:   Heat, Mass and Momentum Transfer;
 Englewood  Cliffs,  New Jersey;  Prentice-Hall, p. 192, Eq. 8-34,  1961.

 Schefer,  R.  W.,  R.  D. Matthews,  N. P. Cernansky, and R.  F.  Sawyer,
 Paper  presented  at  the Western States Section of the Combustion Institute.
 Paper  No.  WSCI 73-31, 1973 Fall  Meeting.

 Schlichting,  H.:   Boundary Layer Theory, Fourth Edition, McGraw-Hill,
 New  York,  1960.

 Seery, D.  J.,  R.  N. Guile, L.  J.  Chiappetta:  Paper No.  14, Presented
 at Eastern States  Section of the Combustion Institute,  Nov. 10-11, 1977.

 Seery, D.  J.  and  M. F. Zabielski:   Private communication,  1979.

 Shapiro,  A.  H.:   The Dynamics  and  Thermodynamics of Compressible Fluid
 Flow,  Vol.  I,  (New  York; Ronald  Press, 1953), p. 228.

 Tine,  G.:   Gas Sampling and Chemical Analysis in Combustion Processes
 Pergamon  Press,  1961.

 Tuttle, J.  H., R. A.  Shisler,  A.  M. Mellor:  "Nitrogen  Dioxide  Formation
 in Gas Turbine Engines:   Measurements and Measurement Methods".   Report
 No.  PURDU-CL-73-06, The Combustion Laboratory, Purdue University, 1973.

 Wilde, K.  A.:   Combustion  and  Flame j_3, p. 173, (1969).

 Vranos, A.,  B. A. Knight,  J. J.  Sangiovanni, L. R.  Boedeker, and D.  J.  Seery:
 Feasibility Testing of Micronized  Coal-Oil (MICO) Fuel  in  a Model Gas  Turbine
 Combustor,  UTRC R79-954451-1,  May  1979.
»D.S. GOVERNMENT FEINTING OFFICE:1981 0-725-402/1111

                                      R-5

-------