paper98...AIAA 98–0712 Reduction and Analysis of Phosphor Thermography Data With the IHEAT...

21
AIAA 98–0712 Reduction and Analysis of Phosphor Thermography Data With the IHEAT Software Package N. Ronald Merski NASA Langley Research Center, Hampton, VA 23681 36th AIAA Aerospace Sciences Meeting and Exhibit January 12–15, 1998/Reno, NV For permission to copy or republish, contact the American Institute of Aeronautics and Astronautics 1801 Alexander Bell Drive, Suite 500, Reston, VA 22091

Transcript of paper98...AIAA 98–0712 Reduction and Analysis of Phosphor Thermography Data With the IHEAT...

Page 1: paper98...AIAA 98–0712 Reduction and Analysis of Phosphor Thermography Data With the IHEAT Software Package N. Ronald Merski NASA Langley Research Center, Hampton, VA 23681 36th

AIAA 98–0712Reduction and Analysis of PhosphorThermography Data With the IHEATSoftware PackageN. Ronald MerskiNASA Langley Research Center, Hampton, VA 23681

36th AIAA Aerospace SciencesMeeting and Exhibit

January 12–15, 1998/Reno, NV

For permission to copy or republish, contact the American Institute of Aeronautics and Astronautics1801 Alexander Bell Drive, Suite 500, Reston, VA 22091

Page 2: paper98...AIAA 98–0712 Reduction and Analysis of Phosphor Thermography Data With the IHEAT Software Package N. Ronald Merski NASA Langley Research Center, Hampton, VA 23681 36th

Reduction and Analysis of Phosphor Thermography Data With the

IHEAT Software Package

N. Ronald Merski�

NASA Langley Research Center, Hampton, VA 23681

Detailed aeroheating information is critical to the successful design of a thermal pro-

tection system (TPS) for an aerospace vehicle. This report describes NASA Langley

Research Center's (LaRC) two-color relative-intensity phosphor thermography method

and the IHEAT software package which is used for the e�cient data reduction and

analysis of the phosphor image data. Development of theory is provided for a new

weighted two-color relative-intensity uorescence theory for quantitatively determining

surface temperatures on hypersonic wind tunnel models; an improved application of theone-dimensional conduction theory for use in determining global heating mappings; and

extrapolation of wind tunnel data to ight surface temperatures. The phosphor method-

ology at LaRC is presented including descriptions of phosphor model fabrication, test

facilities and phosphor video acquisition systems. A discussion of the calibration pro-

cedures, data reduction and data analysis is given. Estimates of the total uncertainties

(with a 95% con�dence level) associated with the phosphor technique are shown to be

approximately 8 to 10 percent in the Langley's 31-Inch Mach 10 Tunnel and 7 to 10percent in the 20-Inch Mach 6 Tunnel. A comparison with thin-�lm measurents using

two-inch radius hemispheres shows the phosphor data to be within 7 percent of thin-�lm

measurements and to agree even better with predictions via a LATCH computational

uid dynamics solution (CFD). Good agreement between phosphor data and LAURA

CFD computations on the forebody of a vertical takeo�/vertical lander con�guration at

four angles of attack is also shown. In addition, a comparison is given between Mach 6

phosphor data and laminar and turbulent solutions generated using the LAURA, GASP

and LATCH CFD codes. Finally, the extrapolation method developed in this report isapplied to the X-34 con�guration with good agreement between the phosphor extrapola-

tion and LAURA ight surface temperature predictions. The phosphor process outlined

in the paper is believed to provide the aerothermodynamic community with a valuable

capability for rapidly obtaining (4 to 5 weeks) detailed heating information needed in

TPS design.

Nomenclature

A Extrapolation factor constantC Heat transfer coe�cient constant, = h(iw=Tw)D Driver constant, = iaw(Tw=iw)� TinitF Flux of light, power/unit areaH Ratio of local heat transfer coe�cient and

stagnation point heat transfer coe�cientI Intensity, projected radiant power

into solid angleK Coating coe�cientL Con�guration lengthM Mach numberPr Prandtl numberQ Radiant Power

�Aerospace Technologist, Aerothermodynamics Branch,

Aero- and Gas-Dynamics Division, Research and Technology

Group, Senior Member AIAA.

The use of trademarks or names of manufacturers in this

paper is for accurate reporting and does not constitute an o�-

cial endorsement, either expressed or implied, of such products

or manufacturers by the National Aeronautics and Space Ad-

ministration.

Copyright c 1997 by the American Institute of Aeronautics

and Astronautics, Inc. No copyright is asserted in the United

States under Title 17, U.S. Code. The U.S. Government has

a royalty-free license to exercise all rights under the copyright

claimed herein for governmental purposes. All other rights are

reserved by the copyright owner.

R Sphere radiusRe Reynolds numberT TemperatureV VelocityX Extrapolation coe�cient

a E�ective aperture factor of camera opticsb Vehicle wing span from wing tip to wing tipc Speci�c heat of model substrateh Heat transfer coe�cienti Enthalpyj Fluorescent light per unit path lengthk Thermal conductivity of model substrate_q Heat transfer rate per unit arear Recovery factort Timex Axial location on con�guration, nose is x=0y Spanwise locationz Distance into substrate

� Weighted logarithmic di�erence

� Variable from Laplace Transform,= Cpt=�

�(T ) Temperature dependent uorescence factor

� Vehicle angle of attack� Thermal product, =

p�ck

1 of 20

American Institute of Aeronautics and Astronautics Paper 98{0712

Page 3: paper98...AIAA 98–0712 Reduction and Analysis of Phosphor Thermography Data With the IHEAT Software Package N. Ronald Merski NASA Langley Research Center, Hampton, VA 23681 36th

Ratio of speci�c heats� Elemental area� Surface emissivity� Ratio of uorscence exponents, �r and �g� Temperature minus initial temperature� Thermal di�usivity, = k=�c� Fluorscence exponent� Running length along con�guration� Density� Stefan-Boltzmann constant� Window transmissivity

Subscripts

FL Flight conditionsG Green component uorescent lightR Red component uorescent lightUV Ultraviolet portion of spectrumWT Wind tunnel conditionsaw Adiabatic wall conditionsconv Convectivee Boundary layer edge conditionseff E�ectivegaw Global adiabatic wall valueinit Initial conditionslam Laminar boundary layer statelamp UV lamprad Radiativerun Wind tunnel run data reduction timest Stagnation pointtl Total conditionstur Turbulent boundary layer statew Conditions at the model surface� Wavelength of light1 Freestream

Introduction

A number of U.S. hypersonic vehicle initiatives haveoccurred over the last few years. For most of theseprograms, the experimental aeroheating data used fordesign of the thermal protection systems (TPS) and es-tablishment of TPS ight margins has been obtained(often exclusively) with the phosphor thermographymethod. The IHEAT phosphor imaging software pack-age was used for the data reduction. This phosphorthermography approach along with the e�cient datareduction by the IHEAT software is revolutionary.Vehicle designers can obtain detailed global, quanti-tative, parametric heating data within a month af-ter a request. Recent applications of phosphor ther-mography and IHEAT include code comparison andcross- ow transition studies on the McDonnell DouglasDelta Clipper Vertical-Takeo�/Vertical-Lander con�g-uration, a shuttle orbiter boundary layer transitionroughness study,1 parametric full con�guration stud-ies in support of Lockheed, Rockwell, and McDon-nell Douglas X-33 Phase 1 concepts, full con�gurationheating studies on the the X-33 Phase 2 concept2, 3

System Monitor

UV Lights

Color CCDCamera

FluorescentModel

Wind TunnelTest Section

Image Processorand Controller

Camera ControlBox

Control Room ComponentsTunnel Components

Workstation

System Monitor

Fig. 1 Phosphor test setup.

and on the X-34 vehicle,4 full con�guration heating onthe X-385{7 and base heating for the Mars Microprobeproject.

Early development of NASA Langley's two-colorrelative-intensity phosphor thermography technique isdescribed in Refs. 8{10. With this method, ceramicwind tunnel models are fabricated and coated withphosphors which uoresce in two regions of the visiblespectrum when illuminated with ultraviolet light. Oneof the phosphors (ZnCdS : Ag;Ni) is a broadbandthermographic phosphor uorescing primarily in thegreen portion of the spectrum and the other is a nar-rowband rare-Earth (La2O2S : 1%Eu) which has twogreen emission spikes and a red emission spike. The uorescence intensity is dependent upon the amountof incident ultraviolet light and the phosphor temper-ature. By acquiring uorescence intensity images witha color video camera of an illuminated phosphor modelexposed to ow in a wind tunnel (see Fig. 1), surfacetemperature mappings are calculated on the portionsof the model which are in the �eld of view of thecamera. This is done by utilizing the green and redcamera outputs and entering the resulting intensityimages into lookup tables created during the calibra-tion of the system. With the present phosphors, ausable temperature range of 70 to 340�F is obtained.In addition, using temperature images acquired at dif-ferent times in a wind tunnel run, heat transfer imagesare computed immediately following a run using theone-dimensional heat conduction equation.

This report presents an overview of the NASA Lan-gley phosphor thermography method and its imple-mentation in the IHEAT data reduction and analysissoftware. It highlights doctoral work which will bepublished in greater detail in Ref. 11. First, a newweighted two-color relative-intensity phosphor ther-mography theory will be introduced which enables agreater degree of accuracy in high-temperature globalphosphor thermography measurements. In addition,an improved application of the one-dimensional heatconduction equation will be presented which increasesthe accuracy of global heating data. Proposed factors

2 of 20

American Institute of Aeronautics and Astronautics Paper 98{0712

Page 4: paper98...AIAA 98–0712 Reduction and Analysis of Phosphor Thermography Data With the IHEAT Software Package N. Ronald Merski NASA Langley Research Center, Hampton, VA 23681 36th

for extrapolating wind tunnel heating data to ightsurface temperatures are also developed. Next, anoverview of the phosphor thermography procedure atthe Langley Research Center will be provided alongwith a description of the individual components of theIHEAT software package. Then, comparisons of phos-phor heating data to thin-�lm measurements and pre-dictions from the LAURA, GASP and LATCH CFDcodes will be presented. Finally an application of theextrapolation method to X-34 laminar and turbulent ight conditions will be given.

Theory

Weighted Two-Color Relative-Intensity

Fluorescence Theory

This section develops a weighted two-color relative-intensity theory for quantitatively determining tem-perature from measurement of uorescent light emit-ted by a model's phosphor coating. In analyzing thisproblem, light from UV lamps illuminates the phos-phor coating, the phosphor coating then emits visiblelight which travels through a wind tunnel window intothe optics of a video camera and �nally to a ChargeCoupled Device (CCD) array. Ultimately, there aretwo primary unknown variables { the local surfacetemperature and the local illumination of UV light onthe model. A number of other factors (including modelcoating characteristics, facility window transmissivity,and optical response among others), however, must beaccounted for or be kept constant from the calibrationprocess to the �nal data reduction.The ux of UV light which excites the phosphor is

equal to the ux of UV light from the lamps times acoating factor.

FUV = KUV Flamp (1)

The coating factor, KUV , is dependent on the ab-sorption rate of the phosphor coating, the coatingthickness, surface re ectivity of the model surface, andincident angle of UV light on a model. A detailed un-derstanding of the dependency of the coating factor onthese variables can be obtained by using the RadiativeTransfer Equation.11

The resulting emission response of the phosphorsfrom the UV illumination is not necessarily linearlyproportional to the amount of UV light incident on themodel. Thus, unless the uorescence characteristicsare the same for two spectral bands or at two di�erenttemperatures, a straight ratio cannot be used to re-move dependence on UV illumination as is often donefor relative-intensity approaches. This is particularlytrue when phosphors are mixed which have di�erentresponse characteristics to UV light. Not accountingfor these uorescence response charateristics can af-fect measurements on wind tunnel models having localUV light illumination levels which signi�cantly deviatefrom intensities used during calibration, particularly

on con�gurations with large amounts of surface cur-vature. This nonlinearity is caused by the phosphorapproaching a saturation limit in regards to excitation.Reference 12 describes a number of possible curve-�tsfor the nonlinear response, the simplest of which isa power curve. The intensity of uorescent light perunit path length emitted by a phosphor illuminatedwith UV light, j�, can therefore be described as:

j� = ��(T )F��(T )UV (2)

All parameters in Equation 2 are dependent upon theregion of the emission spectrum being modelled andthis is denoted by the subscript �.The intensity of light from the surface of the model

can be expressed as the uorescent power per unit pathlength times a coating function:

I� = j�K� (3)

Once the uorescent phosphor light is emitted fromthe model surface, it passes through the tunnel win-dow to the camera optics and reaches the camera CCDarray. The radiant power of light which is incident onthe CCD array can be expressed as:

�Q = �waI (4)

where � is the window transmissivity and a is ane�ective aperture factor which accounts for the trans-mission of light through the camera optics.Substitution of Equations 1 through 3 into Equa-

tion 4 yields

�Q� = �w;�a�K���(T ) [KUV Flamp]��(T ) (5)

Equation 5 primarily depends on the amount ofincident UV light on the model and the local tempera-ture of the phosphor coating. To extract temperaturefrom this equation, the dependence on UV light mustbe eliminated. This is where the relative-intensitymethod comes in. By analyzing the uorescence fromthe model within two di�erent regions of the spectrum,two equations with two unknowns exist and the UVlight dependence can be removed. To do this, the nat-ural logarithm is taken of Equation 5 for two colorcomponents (in this case red and green):

ln(�QR) = ln [�w;RaKR�R(T )]+ �R(T )ln [KUV Flamp](6)

ln(�QG) = ln [�w;GaKG�G(T )]+�G(T )ln [KUV Flamp](7)

At this point, a ratio of the uorescence powers, �is de�ned:

� =�R

�G(8)

Multiplying Equation 7 by 8 and subtracting the re-sult from 6 causes the UV dependence to drop out and

3 of 20

American Institute of Aeronautics and Astronautics Paper 98{0712

Page 5: paper98...AIAA 98–0712 Reduction and Analysis of Phosphor Thermography Data With the IHEAT Software Package N. Ronald Merski NASA Langley Research Center, Hampton, VA 23681 36th

yields the weighted two-color relative-intensity equa-tion:

ln(�QR)� �(T )ln(�QG) = �(T ) (9)

where � is de�ned as the weighted logarithmic di�er-

ence and is given by:

�(T ) = ln [�w;RaKR�R(T )]��(T )ln [�w;GaKG�G(T )](10)

In order to determine the temperature from Equa-tion 9, � and � as a function of temperature mustbe calibrated. Then given incident amounts of greenand red light on the camera array a � value is formedand an iteration has to be performed to back out thetemperature.

Heat Transfer Theory

This section describes the theory used in the re-duction of the phosphor thermography data to heattransfer mappings. In reducing the phosphor data, thephosphor coating is assumed to be in�nitely thin, theconvective heat transfer is assumed to be transmittednormally into the model surface and the local surfaceradius of curvature is assumed to be large. The one-dimensional heat conduction equation can therefore besolved which greatly simpli�es the data reduction. Theone-dimensional approach has its limitations, but forinsulative ceramics such as fused silica and for shorttest times of less than a second, the approximationworks well. In the development of the present heattransfer theory, the thermal properties of the fusedsilica will be assumed to not be dependent on tem-perature. A simple correction to account for thisassumption will be described at the end of the section.The governing one-dimensional heat conduction

equation is given by:

�c@�

@t=

@

@z

�k@�

@z

�(11)

Assuming the thermal properties of the substrate ma-terial to not be dependent on temperature, Equa-tion 11 becomes:

@�

@t= �

@2�

@z2(12)

where � = k=�c is the thermal di�usivity of the sub-strate material.In order to solve Equation 12, one initial condition

and two boundary conditions are required. For an ini-tial condition, the temperature in a direction normalto the surface at a given location is assumed to beconstant before the injection of the model in the windtunnel:

�(z; 0) = 0 (13)

For one of the boundary conditions, an in�nite slabassumption is made whereby the temperature at some

location into the model is always equal to the initialtemperature:

�(1; t) = 0 (14)

For the second boundary condition, the normal con-duction into the wind tunnel model is set equal to theconvective surface heating:

�k@�@zjz=0= _qconv (15)

where _qconv is the convective heat transfer rate perunit area.The convective heating boundary condition is fur-

ther assumed to experience a jump to a constant heattransfer coe�cient during the injection of a model intoa wind tunnel. Typically, for development of simi-lar data reduction approaches, the convective de�ni-tion is temperature based, however, this assumption isstrictly true only for a calorically perfect gas (constantspeci�c heat). In the case of the hypersonic facilities atLangley, the reservoir temperature is past the limits ofa calorically perfect gas and an error can occur whenusing the temperature-based approach. Therefore anenthalpy-based de�nition is used where:

_qconv = h(iaw � iw) (16)

In order to solve the conduction equation usingLaplace transforms, the convective condition needs tobe rewritten in terms of temperature. Pulling out aratio of the surface enthalpy to surface temperature,Equation 16 can be rewritten as:

_qconv = h

�iw

Tw

���Tw

iw

�iaw � Tw

�(17)

For the temperature range over which the phosphorcoating can detect, the air is calorically perfect so theratio of the surface enthalpy to the surface tempera-ture is nearly constant and the convection de�nition isnow in a form which can be used with Laplace trans-forms.Substituting Equation 17 into Equation 15 the sur-

face boundary condition becomes:

�k@�@zjz=0= C(D � �w) (18)

where C and D are constants equal to h(iw=Tw) andiaw(Tw=iw)� Tinit respectively.Equation 12 can now be solved with Laplace trans-

forms using conditions given by Equations 13,14 and18. At the surface of the model, the solution is givenby:

�w

D= 1� e�

2

erfc � (19)

where � = Cpt=�. This equation (which will be re-

ferred to as the step equation) is used for reducing thephosphor data.

4 of 20

American Institute of Aeronautics and Astronautics Paper 98{0712

Page 6: paper98...AIAA 98–0712 Reduction and Analysis of Phosphor Thermography Data With the IHEAT Software Package N. Ronald Merski NASA Langley Research Center, Hampton, VA 23681 36th

0 0.2 0.4 0.6 0.8

(Tw-Tinit)/(Ttl-Tinit)

0

0.25

0.5

0.75

1

1.25

1.5

1.75

2q st

ep/q

code

Temperature-BasedEnthalpy-Based

Fig. 2 Comparison of heat transfer rate compu-

tation between enthalpy-based and temperature-

based methods.

Only an initial temperature and a temperature atsome time into a tunnel run is required to solve theequation. Thus, in a wind tunnel run, an initial tem-perature image can be acquired before a run with no ow on the model and then a temperature image canbe obtained during a run and Equation 19 can besolved at each pixel in the image. No temperature timehistory is needed which is important with an imagingmethod because temperature histories of images wouldrequire a large amount of computer storage space.The e�ect of using an enthalpy-based approach ver-

sus the conventional temperature-based approach isseen in Figure 2. In creating the �gure, a conductioncode was run with a constant heat transfer coe�cientas the surface boundary condition and heat transferrates were computed with both approaches using thesolution's surface temperature value. The computedheat transfer rates were nondimensionalized with theheat transfer rate from the conduction solution. Theresulting heat transfer ratios are plotted versus a tem-perature di�erence ratio, (Tw � Tinit)=(Ttl � Tinit).From the �gure it can be seen that as the wall tem-perature approaches the tunnel total temperature, theerror with the temperature-based method begins toclimb rapidly, while the enthalpy-based method re-mains constant.In reality, the time, t, in Equation 19, does not start

from the moment the model begins to heat. As themodel travels through the test section wall boundarylayer it does not experience a step in the heat transfercoe�cient as was assumed in the heat transfer theory.Therefore, the injection process is modelled as a stepin heating with the step in the heat transfer coe�cientoccurring at some time as the model passes throughthe boundary layer. The e�ective time at which thestep occurs can be calibrated by acquiring data with

a discrete gauge model. The heat transfer rate com-puted using a time-history based method can then beput into Equation 19 and a time backed out. Typicalvariations of the e�ective time on a model are between1 and 2 percent. After performing a number of these\calibrations," a simple approximation for determin-ing the e�ective time has been observed:

teff = trun � twall � (twall � tbl)� 0:5 (20)

where trun is the run time at which the data was ac-quired and tbl is the time that the model reaches theedge of the tunnel boundary layer as determined fromtunnel calibrations. This method of determining thee�ective time is currently the preferred approach. Ef-fectively it exchanges possible inconsistencies in theinjection hydraulics with inconsistencies in the tunnelcalibration pro�le.The thermal product, �, in Equation 19 is deter-

mined from a combination of thermal conductivity andthermal di�usivity measurements from the equation:

� =kp�

(21)

For fused silica, samples have been tested at AnterLaboratories, Inc., of Pittsburgh, PA, to determinetehermal properties and curve-�ts have been computedfrom the resulting data:

� = 9:212605264� 10�6 � 1:91963306� 10�9T

+1:79201� 10�12T 2 m2

s

(22)

k = 0:668170765� 0:000681630� TW

m�K(23)

Up to this point, the heat transfer analysis has as-sumed that the thermal properties, k, �, and � areconstant with respect to temperature although this isnot true. If room temperature properties are assumed,as much as a 17% error can occur at the high endof the phosphor measurement range. For the phos-phor thermography image data reduction, a relativelystraight-forward approach is proposed for correctingthe variable properties e�ects. The thermal productat the initial room temperature, �init, is averaged withthe thermal product at the run temperature, �run, togive an e�ective thermal product, �eff , used for datareduction. Thus,

�eff =�init + �run

2(24)

To check this equation, a one-dimensional in�nite slabconduction code was run in a variable property modefor a range of heat transfer coe�cients. The code wasrun for both fused silica and Macor thermal proper-ties and resulting surface temperatures were reduced

5 of 20

American Institute of Aeronautics and Astronautics Paper 98{0712

Page 7: paper98...AIAA 98–0712 Reduction and Analysis of Phosphor Thermography Data With the IHEAT Software Package N. Ronald Merski NASA Langley Research Center, Hampton, VA 23681 36th

0 0.2 0.4 0.6 0.8

(Tw-Tinit)/(Ttl-Tinit)

0.95

0.96

0.97

0.98

0.99

1

1.01

1.02

1.03

1.04

1.05q st

ep/q

code

Fused SilicaMacor

Fig. 3 Error of variable thermal properties ap-

proximation.

using Equation 19. The results are plotted in Fig-ure 3. Here again, the ratio of reduced heat transferrate to the heat transfer rate from the code, qstep=qcode,is plotted against the temperature di�erence ratio,(Tw � Tinit)=(Ttl � Tinit). The heat transfer ratesfor both the corrected fused silica and the Macor arewithin 0.4% for the full range of temperature di�erenceratios. This error is less than the error associated withabsolute values of the thermal properties, so Equa-tion 24 is believed to be su�cient for data reductionwith fused silica and Macor substrates.

In order to determine an adiabatic wall enthalpy,iaw,the ratio of the adiabatic wall enthalpy to the totalenthalpy can be found from the expression:

iaw

itl=

1 + r �12M2

e

1 + �12M2

e

(25)

For laminar ows, the recovery factor, r, is equaltopPr and for turbulent ow it is equal to Pr1=3.

The edge Mach number can be estimated by usingtangent-wedge approximations with the oblique shockrelations. Along the body, the adiabatic wall enthalpydecreases from the total enthalpy. Currently for thephosphor thermography method the local surface an-gle at a pixel is not known. So with laminar extrap-olations, the strategy is to �nd the lowest adiabaticwall enthalpy on a con�guration and then averagethis result with the total enthalpy to obtain a globaladiabatic wall enthalpy. In the case of turbulent ex-trapolations where the ow trips back on the body, theadiabatic wall enthalpy is just selected as the adiabaticwall enthalpy on a wedge with an angle equal to theangle of attack of the vehicle.

Extrapolation of Phosphor Data To Flight Surface

Temperatures

Once the heat transfer coe�cient mappings havebeen determined via wind tunnel tests, they can beused to validate CFD predictions, to determine dis-persion heating factors or to predict the surface tem-perature of a vehicle in ight. This last use for thedata, extrapolation to ight, has the possibility of sav-ing large amounts of computational time required for ight predictions. This section describes a methodwhich has been developed for use with IHEAT phos-phor images and has been successfully applied to theX-34 con�guration.A standard approach to extrapolationg wind tunnel

aeroheating data to ight is to reference heat transfercoe�cients on a scaled con�guration in the wind tun-nel to the heat transfer coe�cient at the stagnationpoint on a reference hemisphere at the tunnel condi-tions. As described in Ref. 13, this ratio can thenbe related to the heat transfer coe�cient for the fullcon�guration in ight ratioed by the heat transfer co-e�cient of a scaled up reference hemisphere using theexpression:

HFL

HWT

= X (26)

where the H 's are the ratio of the local surface heattransfer coe�cient divided by the stagnation pointheat transfer coe�cient on a sphere at ight or windtunnel conditions and X is an extrapolation factor.The extrapolation factor adjusts the data for di�er-ences between wind tunnel and ight such as in Machnumber, Reynolds number and the existence of turbu-lence.Heat transfer rate correlations reported by Tauber

and Meneses14, 15 were used as a guide in developingthe form for the extrapolation factors. The local heattransfer coe�cient at a sphere stagnation point is as-sumed to be approximated by:

hst = �:51V 31

�R

L

��:5

L�:5Ast

itl(27)

where R is the radius of the sphere, L is the length ofthe geometry and A is a constant dependent on angleof attack and geometry considerations. Likewise for alaminar at plate:

hlam = �:51V 3:21

��

L

��:5

L�:5Alam

itl(28)

where � is the running length from the start of astreamline. Finally in the case of turbulent heatingon a at plate:

htur = �:81V 3:71

��

L

��:2

L�:2Tw�:25Atur

itl(29)

where � in this case is the running length from thestart of transition.

6 of 20

American Institute of Aeronautics and Astronautics Paper 98{0712

Page 8: paper98...AIAA 98–0712 Reduction and Analysis of Phosphor Thermography Data With the IHEAT Software Package N. Ronald Merski NASA Langley Research Center, Hampton, VA 23681 36th

Extrapolation factors for laminar and turbulent ows can now be determined by dividing the local lam-inar or turbulent h (Equations 28 and 29 respectively)by the stagnation point h (Equation 27). The resultingratio at ight conditions is then divided by the ratio atwind tunnel conditions. After cancellations are madefor the running lengths, �, and the constants, A, thefollowing laminar factor results:

Xlam =Hlam;FL

Hlam;WT

=

�V1;FL

V1;WT

�:2

(30)

and in the case of turbulent ow

Xtur =Htur;FL

Htur;WT

=

��1;FL

�1;WT

�:3�V1;FL

V1;WT

�:37

��LFL

LWT

�:3�Tw;FL

Tw;WT

��:25

(31)

The heating at a point on the ight con�gurationis then determined by calculating the stagnation pointheat transfer coe�cient at a point on the trajectory ona sphere at the ight scale. The ratio from the windtunnel is multiplied by this ight stagnation point co-e�cient and the extrapolation factor, X , to yield the ight heat transfer coe�cient. The heat transfer rateis determined by multiplying the heat transfer coe�-cient by the thermal driver, (iaw � iw).The wall temperature in ight is predicted by as-

suming that the radiative heat transfer rate at thesurface of the ight vehicle, _qrad, is equal to the convec-tive heat transfer rate into the vehicle surface, _qconv:

_qrad = _qconv (32)

From the Stefan-Boltzmann law, the radiative heattransfer rate is related to the surface temperature,TFL, by:

_qrad = ��T 4FL (33)

where � is the surface emissivity and � is the Stefan-Boltzmann constant. Substituting Equation 33 andthe de�nition of the heat transfer coe�cient into Equa-tion 32 yields:

_qrad = ��T 4FL = hFL(iaw;FL � iw;FL) (34)

Since the wall enthalpy is dependent on the surfacetemparature, Equation 34 must be solved iteratively.If the ight total enthalpy is high enough however, thewall enthalpy can often be assumed to be small as com-pared to the adiabatic wall enthalpy and Equation 34becomes

_qrad = ��T 4FL = HFL iaw;FL (35)

This \cold wall" assumption, when applicable, re-moves the need for the iteration associated with Equa-tion 34 and greatly simpli�es the computations re-quired for global images.

The adiabatic wall enthalpy is a fraction of the total ight enthalpy. To calculate the total ight enthalpy,half the square of the velocity is added to the localambient enthalpy:

itl = V 21=2 + i1 (36)

The enthalpy at the wall is determined using the ther-modynamic property curve-�ts for individual speciesreported in Ref. 16. Knowing the species fractionsof air at a temperature and altitude, the enthalpycan be calculated. For this analysis, sea-level volumefractions of N2, O2, and Ar as de�ned by the U.S.

Standard Atmosphere, 1976 17 are used. Even thoughthe number densities of these species drop as the alti-tude increases, the fractional volumes remain constantup to 282,000ft: and so the same enthalpy formulationcan be used up to this altitude. At higher altitudesthough, monatomic oxygen becomes a signi�cant com-ponent of the ambient air and the fractional volumesmust then be computed for each altitude.As the temperature increases to 3000�F , the air is

no longer thermally perfect. The oxygen begins todissociate and the proportions of the constituents ofthe air changes. Properties for air under equilibriumconditions can be determined, but a pressure depen-dency arises in the computation of the enthalpy. Forthis global extrapolation method, the surface pres-sure is not known on the vehicle so the method isthus restricted to surface temperatures of less than3000�F if a cold-wall assumption cannot be applied, al-though typically when these temperatures are reachedthe cold-wall assumption is valid.Upon extrapolating to ight, the adiabatic wall en-

thalpy is chosen to be the same fraction of the totalenthalpy as in the case of the tunnel conditions, thusproviding a consistent convective heat transfer de�-nition. There is still an error in the computed heattransfer rate, however, resulting from the deviation ofthe chosen global adiabatic wall value from the trueadiabatic wall value (proof not shown) of

_qFL;gaw_qFL;aw

=

�iaw;WT � iw;WT

igaw;WT � iw;WT

��igaw;FL � iw;FL

iaw;FL � iw;FL

(37)where the subscript, gaw, stands for the global adia-batic wall value chosen for the whole phosphor image.

Phosphor Thermography Method in

Langley Facilities

Model Fabrication

Fabrication of wind tunnel models for phosphor test-ing is a critical component in the technique. In orderto obtain accurate heat transfer data using the one-dimensional heat conduction equation, models need tobe made of a material with a low thermal di�usivityand well-de�ned, uniform, isotropic thermal proper-ties. Also, the models must be durable for repeated

7 of 20

American Institute of Aeronautics and Astronautics Paper 98{0712

Page 9: paper98...AIAA 98–0712 Reduction and Analysis of Phosphor Thermography Data With the IHEAT Software Package N. Ronald Merski NASA Langley Research Center, Hampton, VA 23681 36th

use in the wind tunnel and they should be subject tominimal deformation when thermally cycled. To meetthese requirements, Langley has developed a unique,silica ceramic slip casting method.18 Patterns for themodels are typically made using a numerical cuttingmachine or with the stereo-lithography process. Usingthese patterns, investment molds are created. Ceramicslip is poured into the molds and after the ceramicsets, the investment molds crumble o� leaving ceramicshells. These shells are next �red in a kiln and pot-ted with support hardware. The slip casting methodallows for the casting of relatively �ne details as wellas thin ceramic sections such as wings and �ns. Modellengths are typically 10 to 14 inches; however, mod-els up to 30 inches in length have been fabricated fortesting at low angles of attack. In addition, the slip-casting method is a rapid process whereby, in three tofour weeks, a full array of models can be fabricated,complete with various perturbations in a con�gurationsuch as nose radius changes and control surface de ec-tions.

Once the models are fabricated, they must be coatedwith phosphor crystals. Currently, the phosphors aresuspended in a silica ceramic binder and applied withan airbrush. Final coating thicknesses have been mea-sured to be approximately .001 inches. The coatingmethod which has been developed, produces robustcoatings which do not require reapplication betweenruns, thereby signi�cantly enhancing the e�ciency ofthe phosphor technique.

Facilities

Langley's phosphor thermography method has beenused in four of the facilites in NASA Langley'sAerothermodynamic Facilities Complex (AFC): the31-Inch Mach 10 Air, 20-Inch Mach 6 Air, 20-InchMach 6 CF4 and 15-Inch Mach 6 High TemperatureTunnels. A detailed description of these facilities isgiven by Micol.19 This section brie y describes the�rst two facilities where the phosphor method has beenprimarily used to date.

The 31-Inch Mach 10 facility consists of high pres-sure air storage rated to a maximum pressure of 4400psia, a 12.5 MW electrical bundle heater which heatsthe air to approximately 1850 �R, a settling chamber,three-dimensional contoured nozzle, test section, ad-justable second minimum, aftercooler, vacuum spheresand vacuum pumps. A 5-micron in-line �lter, beforethe nozzle, removes particles from the air- ow whichcan damage the ceramic phosphor models. The noz-zle itself has a 1.07-in. square beryllium copper throatand its exit height is 31 inches. The test section hasa 30-inch by 17-inch tempered-glass window for illu-minating and viewing the phosphor models. Windtunnel models are protected from the ow as the tun-nel comes to its operating condition by a door in theside-wall of the test section. After the operating con-

dition is reached, the model is injected to the tunnelcenterline in approximately 0.55 sec from the back sideof the test section with a hydraulic injection system.Freestream Reynolds numbers obtained in this tunnelvarying from 0.25 to 2.2�106=ft.The 20-Inch Mach 6 Tunnel has a two-dimensional

contoured nozzle, with the bottom and top walls be-ing contoured and the side walls parallel. The testsection size is 20.5 in: by 20 in: and has temperedglass windows on the top and two sides. Models aremounted on an injection system in a housing belowthe test section. During a run, the models can be in-jected in less than 0.5 sec: The tunnel has a range ofoperating reservoir pressures from 30 to 500 psia andreservoir temperatures of up to 1000 �R. This corre-sponds to a freestream Reynolds number range of 0.5to 9.0�106=ft.Acquisition Systems

Four phosphor thermography video acquisition sys-tems have been constructed for use in Langley's hyper-sonic facilities, two of which are currently operationaland two which will be brought online in the near fu-ture. The systems are composed of a Sony XC007video camera with a zoom lens and a PC computerequipped with a Matrox Meteor digitizer. The SonyXC007 camera has 3 CCD arrays which obtain red,green and blue component images (only red and greenbeing used for phosphor thermography). Each arrayhas 768 horizontal picture elements and 493 verticalpicture elements. Signal from the color CCD arraystravels into a camera control box which performs someanalog signal processing. Processing features such as agamma compensation circuit and a linear matrix cir-cuit for obtaining \life-like" color are disabled.Once the signal has been obtained and processed by

the camera control box, it then proceeds to the Ma-trox Meteor card in the PC where the video frames aregrabbed and digitized with an analog-to-digital con-verter (ADC). The ADC digitizes the signal to 8 bitswhich corresponds to 256 discrete counts. The wayin which the signal can be digitized can be modi�edby black and white level inputs to the ADC from thesystem software.The acquisition software makes use of the Matrox

Imaging Library routines. Digital interlaced imagescan be obtained at the standard video rate of 30 framesper second with a spatial resolution of 640 by 480 pix-els. The system can be put into a live video modewith a pseudo-color color lookup table to set up theUV lighting on the models.

Tunnel Tests

In order to set up for a wind tunnel run, the �rst taskafter mounting a model is to set up the UV lighting.The model is injected into the tunnel and a monitoris set next to the test section which displays a livepseudo-color image of the green component from the

8 of 20

American Institute of Aeronautics and Astronautics Paper 98{0712

Page 10: paper98...AIAA 98–0712 Reduction and Analysis of Phosphor Thermography Data With the IHEAT Software Package N. Ronald Merski NASA Langley Research Center, Hampton, VA 23681 36th

Fig. 4 Typical IHEAT interface window.

video camera. The lighting is set so that the uo-rescence intensity from the model does not go overthe scale limits of the phosphor system. This pro-cess is performed on the green component because thiscomponent decreases throughout a run and ultimatelyquenches out if a high enough temperature is reached.By keeping the green uorescence intensity as high aspossible without going overscale, temperature rangeand resolution is optimized.Once the lighting is set, the model is retracted and is

ready for the tunnel run. The acquisition system is setto acquire images at di�erent times into the run. Dueto possible conduction e�ects, earlier images providedata on high-heating, high-curvature portions of themodel and later images provide data for low-heating,low-curvature portions of the model. After the acqui-sition system and tunnel are ready to run, the modelis brie y injected into the tunnel test section and apre-run image is obtained of the initial temperaturedistribution on the model. The model is retracted andthe tunnel is brought up to the desired reservoir pres-sure and temperature conditions. The model is theninjected, the acquisition system is triggered and im-ages are acquired at the preset times. Once the runhas been completed, the images are saved and portedover to a UNIX-based workstation for data reductionand analysis.

Phosphor Calibration, Data Reduction

and Analysis

IHEAT

The IHEAT software package is a UNIX-based,\user-friendly" software suite of eight programs fore�ciently handling the large amounts of image data as-sociated with the phosphor thermography calibration,data reduction and data analysis. A typical IHEATmain window, shown in Fig. 4, contains an image area,a menu bar at the top and a set of tool buttons alongthe left side. The menu bar includes menu buttons forloading in �les, working with image titles, color barranges, color tables, data processing and more. All

REGRID EXTRAP DISPER

Data Analysis Programs

ExtrapolatedData

Extrapolation/CFD Comparisons

IHEAT Products

YawDispersions

TunnelRepeatability

Wind Tunnel Data/CFD Comparisons

Wind TunnelAeroheating Data

Reynolds No.Changes

Calibration Routines

SYSCAL TRANSCAL

TEMPCAL LUTCALC

CFD ResultsIHEAT Data

Reduction

Acquisition Systems

Fig. 5 The IHEAT software environment.

processes in IHEAT are performed using the metricsystem, but a menu button exists for displaying re-sults in English units. A set of tool buttons on the leftside includes three options for extracting data fromthe IHEAT images with line cuts, a statistical tool foranalyzing regions of interest in an image, a zoom fea-ture, a button to mask over spurious data in an imageand two buttons for arrow and text annotation.

Figure 5 shows a schematic of the IHEAT soft-ware environment. Calibration or wind tunnel imagedata are acquired by the acquisition systems and sentto the calibration programs or the main data reduc-tion program. Three calibration programs, SYSCAL,TRANSCAL, and TEMPCAL are used for reducingcalibration data and the LUTCALC program packagesthe calibration results into temperature lookup tablesfor use by the IHEAT data reduction program. Oncethe phosphor data has been reduced, three programsexist for use in further analyzing the results. The DIS-PER program overlays images from di�erent runs andprovides ratios of heating between runs. EXTRAPuses the extrapolation theory previously described todetermine laminar and turbulent heating levels at apoint on a trajectory. The REGRID program is aninterpolation program where CFD surface data can bepackaged together with phosphor mappings to makedirect comparisons in the global characteristics be-tween the computations and the wind tunnel data.

Calibration

The phosphor thermography method has been de-veloped to support work in a number of hypersonicfacilities. In addition, four acquisition systems havebeen constructed for utilization in the facilities. Theideal situation would be a calibration process for onesystem, at one facility and for one phosphor mix. How-ever with the number of systems involved, with theway acquisition systems can be shuttled between fa-cilities and with the batch-to-batch variations which

9 of 20

American Institute of Aeronautics and Astronautics Paper 98{0712

Page 11: paper98...AIAA 98–0712 Reduction and Analysis of Phosphor Thermography Data With the IHEAT Software Package N. Ronald Merski NASA Langley Research Center, Hampton, VA 23681 36th

0 50 100

Intensity (candela/m2)

0

50

100

150

200

250S

yste

mR

esp

onse

(dig

italc

oun

ts) Red response data

Red response fitRed response (repeat)

Fig. 6 Red component system calibration.

can occur with the phosphors, the matrix of requiredcalibrations can rapidly expand.

In order to reduce the number of calibrations whichare required, a modular approach has been developed.Each system is calibrated individually for system re-sponse, each facility window is calibrated for transmis-sivity in the red and green portions of the spectrum,and the temperature/intensity response characteristicsfor each phosphor batch are separately calibrated. Allof the calibrations are �nally packaged together forany given combination via a lookup table creation pro-gram.

The �rst calibration required is to determine the re-sponse of the system to a given input of light, �Q,for use in Equation 9. To perform this, an acquisi-tion system is pointed at an integrating sphere thatis stepped through di�erent intensity levels. Imagesare acquired at each level. In reducing the calibrationdata, the SYSCAL program is used. A region of in-terest is picked for the range of images and analyzed.From this analysis, a fourth-order �t is made for eachcolor component of digital system response in countsversus light intensity. A sample red component cali-bration, along with the �t and a repeat run, is shownin Figure 6.

In characterizing the facility window transmissivi-ties, an integrating sphere is placed behind a windowat an angle relative to the window. On the other sideof the window, a camera is set up and images are ac-quired of the integrating sphere port, with and withoutthe window between the port and the camera. Redcomponent and green component images are broughtinto the TRANSCAL program and converted to lightintensity mappings and the resulting mappings are av-eraged to yield window transmissivity factors in boththe red and green portions of the spectrum.

The �nal calibration determines the uorescence re-

0 1 2 3 4 5

ln(δQg)

0.5

1

1.5

2

2.5

3

3.5

4

4.5

5

ln(δ

Qr)

Increasing Temperature

20°C

170°C

Fig. 7 Sweep of log-log lines.

50 100 150 200 250 300 350

Temperature (° F)

0

0.5

1

1.5

ηTemperature IncreasingTemperature DecreasingCurve-Fit

Fig. 8 Ratios of uorescent exponents.

sponse characteristics of the phosphors at di�erenttemperatures. In performing this calibration, a thin3-inch square plate coated with phosphors is placedinto a computerized convection oven which has a win-dow on the front. The plate is illuminated with a UVlamp from outside the oven with a range of incidentintensities. During a calibration the plate is viewedwith an acquisition system and images are acquiredat temperatures ranging from 70 to 340 �F . Afterthe data is acquired, the images are analyzed with theTEMPCAL program. At each temperature, log(�Qr)-log(�Qg) lines are determined with slopes of �(T ) andy-intercepts of �(T ) as described by Equation 9. Asample sweep of lines from a calibration is shown inFigure 7. Once the slopes and intercepts have beendetermined, they are plotted versus temperature asshown in Figures 8 and 9.Lookup tables are created from the calibrations for

10 of 20

American Institute of Aeronautics and Astronautics Paper 98{0712

Page 12: paper98...AIAA 98–0712 Reduction and Analysis of Phosphor Thermography Data With the IHEAT Software Package N. Ronald Merski NASA Langley Research Center, Hampton, VA 23681 36th

50 100 150 200 250 300 350

Temperature (° F)

0

0.5

1

1.5

2

2.5∆

Temperature IncreasingTemperature DecreasingCurve-Fit

Fig. 9 Logarithmic di�erences.

use in IHEAT by running the LUTCALC program.When using LUTCALC, variables are set correspond-ing to a desired combination of the three calibrationsand LUTCALC iterates on each of the possible redand green acquisition readings, ultimately creating a256 by 256 temperature lookup array �le.

Data Reduction

Data reduction is performed in the IHEAT program.When reducing the data, a setup �le is �rst loadedinto IHEAT which contains all of the speci�c informa-tion for the test setup including phosphor system used,facility that the test is being performed in, and thephosphor mix which has been applied to the models.Once the setup �le has been loaded, a �le with the tun-nel ow conditions for the run being analyzed is input.This �le includes the tunnel total temperature, modelangle of attack and a predicted stagnation point heat-ing value which IHEAT uses to non-dimensionalize thedata. Next, the position history of the model is inputand analyzed to precisely determine the e�ective timeat which the images are to be reduced. The pre-runintensity images and run intensity images are inputinto IHEAT and immediately converted to tempera-ture mappings.Once the �les have been loaded, the next step is

to compute heat transfer coe�cients. If a global adia-batic wall temperature is desired other than the tunneltotal temperature, a menu buttion exists which allowsfor the calculation of the adiabatic wall temperature ona wedge with a half angle corresponding to the modelangle of attack using Equation 25. At this point, theglobal adiabatic wall temperature is known, along withthe e�ective time and the model run and pre-run tem-peratures. Equation 19 is solved at every point on themodel which is in the �eld of view of the camera.After the heat transfer mappings have been com-

puted, often line cuts from the image are desired. An

automated approach exists in IHEAT to do this. De-sired axial and longitudinal cut locations are enteredinto a window in IHEAT. Next, known �ducial marklocations in the image are clicked on and also inputinto IHEAT. IHEAT cuts the images along the desiredcut locations, stretching the cuts according to the in-put �ducial mark locations. Currently, the full surfacegeometry is not entered into IHEAT, so, more �ducialmarks input into IHEAT corresponds to a greater ac-curacy in the locations of the cut data. Once the cutshave been obtained, they are automatically saved forcomparison with cuts from other runs.

Data Analysis

IHEAT data can be extrapolated to ight surfacetemperature levels using the EXTRAP program. Todo this, a reduced IHEAT �le is entered into EXTRAPalong with the desired trajectory point (described bythe vehicle velocity and altitude). In addition, thepredicted heat transfer coe�cient at ight conditionsmust be input for a hemisphere of a radius scaled upfrom that of the hemisphere which was used to nondi-mensionalize the tunnel data. The type of correctionfor the state of the boundary layer (laminar or turbu-lent) must also be entered and, if it is valid, a cold-wallassumption toggle can be set to reduce computationtime. Once the parameters are set, EXTRAP then cal-culates the surface temperature mapping at the ightconditions using Equations 30 through 36 as appro-priate. After the surface temperature image has beendetermined, temperature line cuts can be extracted o�of the image in the same way as with the tunnel data.

The second analysis program available to IHEAT isthe DISPER program. Two reduced IHEAT runs areentered into this program, a baseline and some dis-persion such as in yaw. The DISPER program helpsthe user in registering the two images over each otherand then divides the dispersion run by the baseline.A number of possible uses for this analysis exist in-cluding examination of slight changes in yaw, angleof attack e�ects, heating augmentations on de ected aps, Reynolds number variations, run-to-run tunnelrepeatability, comparison of results between facilitiesand substrate conduction error estimation.

The �nal IHEAT analysis code is the REGRID pro-gram. REGRID is a program which converts eitherCFD ight surface temperature predictions or windtunnel heat transfer rate predictions to images andputs the resulting images side by side with phosphordata. This is done by splitting the CFD surface gridcomputational cells into triangular elements, interpo-lating within each element and projecting the resultback into the phosphor imaging space. The phosphorand CFD images are then displayed together alongthe symmetry plane of the geometry. Currently, thiscomparison can be performed for lee and wind viewsonly. Since the surface geometry is available from the

11 of 20

American Institute of Aeronautics and Astronautics Paper 98{0712

Page 13: paper98...AIAA 98–0712 Reduction and Analysis of Phosphor Thermography Data With the IHEAT Software Package N. Ronald Merski NASA Langley Research Center, Hampton, VA 23681 36th

computational grid, the phosphor image has also beenstretched in the past to account for curvature e�ectswhen not enough �ducial marks were placed on themodel to adequately de�ne a geometry.

Error Analysis

In analyzing the error for the phosphor thermogra-phy method, the AIAA standard for assessing windtunnel data uncertainty20 was used which in turn con-forms with international guidelines and standards.21

While a number of di�erent uncertainty sources ex-ist for phosphor data including those due to calibra-tion, data acquisition and processing, lack of �delityin the models, knowledge of the tunnel test environ-ment, three-dimensional conduction e�ects within thesubstrate, and registration of image pixel location tophysical location on the model, this analysis reportserrors associated only with the �rst two sources (cal-ibration and data acquisition and processing) whichare the most readily quanti�able. Errors are pre-sented with bias limits, precision limits and overalluncertainty (which is the root of the sum of the un-certainties squared). All errors are reported with 95%con�dence limits. The analysis describes the uncer-tainty associated with each variable in Equation 19(initial temperature, run temperature, adiabatic walltemperature, substrate thermal properties and e�ec-tive time) and then presents the combined uncertaintyfor all of the variables.Temperature measurement uncertainties are plotted

in Figure 10. The bias error was determined from theuncertainty in the calibration oven temperature read-ing. The temperature distribution error through theoven test section was quoted to be 0.026 �F and wasconsidered negligible. The oven thermocouple uncer-tainty (furnished by the manufacturer) varied from 3.6to 4.9 �F . The precision error was found by puttingthe temperature calibration images (which theoret-ically are isothermal) into IHEAT and statisticallylooking at the temperature measured at each pixelelement, calculating a standard deviation from thenominal temperature reading in the calibration andmultiplying by two for the 95% limits. From Figure 10the temperature error varies from 5.2 to 6.5�F . It isseen that the precision uncertainty dominates at thelower temperature, but then the bias uncertainty dom-inates the overall uncertainty at temperatures higherthan 140�F . For the combined uncertainty analysis,the error in the initial temperature was assumed to be5.2�F and the wall temperature error was assumed tobe at the worst case of 6.5�F through the temperaturerange.In order to compute the uncertainty in the adiabatic

wall enthalpy, a worst case situation was applied. Theadiabatic wall enthalpy for a plate at zero degrees angleof attack was calculated and averaged with the tunneltotal enthalpy. The di�erence between the averaged

0 100 200 300

Temperature (° F)

0

1

2

3

4

5

6

7

8

9

10

Err

or(°

F)

Precision UncertaintyBias UncertaintyOverall Uncertainty

Fig. 10 Temperature errors.

value and the tunnel total enthalpy was used as theerror.

For the substrate thermal properties, the quotedinstrumentation uncertainty was 3% for both the dif-fusivity and conductivity and this was used as a biasuncertainty. A statistical analysis with a number ofsamples yielded precision uncertainties of 9.6% and1.5% for di�usivity and conductivity with correspond-ing overall uncertainties of 10% and 3.4%. The errordue to choosing an average thermal product to accountfor thermal properties variations (Equation 24) waspreviously found to be 0.4% and was assumed to beinsigni�cant. While the di�usivity error seems large,the heat transfer rate is not ultimately as sensitive toit since a square root is taken of the di�usivity withinthe data reduction.

The e�ective time from Equation 20 was found fromthin-�lm calibration runs to have an uncertainty of0.055 sec. At 1 second, the error is thus 5.5%.

In determining a combined uncertainty, a standardapproach is to use a �rst order Taylor series expansionwhere sensitivities of variables within an equation arefound by taking the partial derivative of the equationwith respect to each variable within the equation. Thisapproach is only valid, however, when a linear functionis being analyzed. The function in Equation 19 is notlinear and when a �rst order Taylor series is used, ar-ti�cially large errors arise due to the contribution ofthe adiabatic wall enthalpy. In order to properly com-pute the errors, higher order terms in the Taylor seriesexpansion must be used. This approach became im-practical, so a computer program was written whereeach of the variables in Equation 19 was varied by itsuncertainty and the resulting heat transfer rate wascompared with the heat transfer rate calculated withbaseline values for the variables.

Combined uncertainties in heat transfer rate for the

12 of 20

American Institute of Aeronautics and Astronautics Paper 98{0712

Page 14: paper98...AIAA 98–0712 Reduction and Analysis of Phosphor Thermography Data With the IHEAT Software Package N. Ronald Merski NASA Langley Research Center, Hampton, VA 23681 36th

100 200 300

Surface Temperature (°F)

0

2

4

6

8

10

12

14

16

18

20U

nce

rtai

nty

(95%

conf

iden

celim

its) Thermal Properties

TimeInitial TemperatureSurface TemperatureDriver EnthalpyTotal

Fig. 11 Heat transfer rate errors in 31-Inch Mach

10 Tunnel.

100 200 300

Surface Temperature (°F)

0

2

4

6

8

10

12

14

16

18

20

Un

cert

aint

y(9

5%co

nfid

ence

limits

) Thermal PropertiesTimeInitial TemperatureSurface TemperatureDriver EnthalpyTotal

Fig. 12 Heat transfer rate errors in 20-Inch Mach

6 Tunnel.

phosphor data versus measured model surface temper-ature at 1 second into a run are plotted for the 31-InchMach 10 Tunnel and 20-Inch Mach 6 Tunnel in Fig-ures 11 and 12 respectively. From the plots, it canbe observed that the initial temperature and run tem-perature are the dominant uncertainties when lowersurface temperatures are measured. Indeed, as the sur-face temperature approaches the initial temperature,the uncertainty climbs to in�nity. This high error dueto temperature is responsible for the large amount ofscatter observed in phosphor heating data obtainedfrom lower surface temperature readings. In mostcases though, the wall temperature measurements arehigher and the uncertainty due to thermal properties isthe signi�cant error source. In the case of the 20-InchMach 6 tunnel, at the higher end of the temperature

Fig. 13 Two-inch radius thin-�lm hemisphere.

range, the surface temperature to adiabatic wall tem-perature is much larger (due to the lower reservoirtemperature of the tunnel) and the contribution of theerror in the adiabatic wall temperature is larger thanin the case of the 31-Inch Mach 10 Tunnel. For themajority of the temperature range, the resulting totalphosphor uncertainty varies between 7 to 10% in the31-Inch Mach 10 tunnel and 8 to 10% in the 20-InchMach 6 tunnel.

Comparison to Thin-Film Hemisphere

Measurements

A 2-inch radius phosphor hemisphere model was fab-ricated and tested to compare with measurements ona 2-inch radius hemisphere instrumented with thin-�lms. The phosphor hemisphere was made using thefabrication approach previously described. A small�ducial mark was placed at the stagnation point in or-der to determine the stagnation point location in therun images. In the case of the thin-�lm hemisphere,the model (shown in Figure 13) was made from Ma-cor since the �lms could not be successfully appliedto a fused slica substrate. Nine gauges were placedalong an arc at two and a half degree intervals fromthe stagnation point back to the twenty degree loca-tion. A gauge was placed at 25 degrees and then from30 to 70 degrees in 10-degree increments. Four ad-ditional gauges were placed in two and a half degreeintervals back from the stagnation point on the otherside of the model as a check on the symmetry of themeasurements.Both models were tested in the 31-Inch Mach 10

Tunnel at reservoir conditions of 720 psi and 1830�R,corresponding to a freestream Reynolds number of1.0x106/ft. The phosphor data was reduced at time of1 second after the model had been exposed to the ow.The thin-�lm measurements were reduced with a datareduction code developed by Hollis22 which uses theKendall-Dixon-Schulte23 one-dimensional time historyapproach. All data were non-dimensionalized usingthe Fay-Riddell24 stagnation point heating values cal-culated using the modi�ed Newtonian method for thepressure gradient.

13 of 20

American Institute of Aeronautics and Astronautics Paper 98{0712

Page 15: paper98...AIAA 98–0712 Reduction and Analysis of Phosphor Thermography Data With the IHEAT Software Package N. Ronald Merski NASA Langley Research Center, Hampton, VA 23681 36th

0 20 40 60 80

Angle (degrees)

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

1.1h/

h FR

Thin-film dataCFD (LATCH)Phosphor data

Fig. 14 Two-inch radius hemisphere

phosphor/thin-�lm comparison.

CFD predictions were also made using the LATCHprogram for comparison with the wind tunnel data.LATCH uses an axisymmetric analog for three-dimensional boundary layers with boundary layeredge conditions from an inviscid ow�eld solution.25

LATCH computations for this comparison were per-formed, assuming equilibrium air properties, at windtunnel conditions and non-dimensionalized in the sameway as the wind tunnel data.

A cut on the phosphor image was taken from thestagnation point to the aft end of the sphere. Thephosphor data from the 80 to 90 degree locations wasremoved due to low UV illumination which put thedata out of the phosphor calibration range. Resultsare plotted along with the thin-�lm measurements andthe LATCH predictions in Figure 14. Two error barshave been placed on the plot, one near the front ofthe sphere with an uncertainty of �7.5% and one nearthe rear of the plot with an uncertainty of �15%. Thedata and LATCH predictions fall within a 7 percentband through the full 80 degrees plotted in the �gure.Through the �rst 50 degrees, however, the phosphordata tends to agree better with the CFD computationsthan with the thin-�lm data. The stagnation pointheating on a hemisphere is expected to be higher thanFay-Riddell theory when the Fay-Riddell calculationis performed using the modi�ed Newtonian methodto determine the pressure gradient. This implies thatthe thin-�lm heat transfer data may be low in thestagnation point region for reasons as yet to be de-termined. Based on this comparison, however, thephosphor heating levels agree with thin-�lm measure-ments and computations within the uncertainties ofthe methods, thereby providing a degree of con�dencein the phosphor technique.

Computations

Experimental Data

0.500.00 0.75 1.000.25h/hFR

Fig. 15 VT/VL Mach 10 Phosphor/LAURA com-

parison, � = 17.5�.

Computational Comparisons

Mach 10 8-Degree Sphere Cone Comparison

In support of the competitive �rst phase of the X-33 program, phosphor measurements were performedon a vertical takeo�/vertical lander (VT/VL) con�g-uration and compared with CFD predictions. TheVT/VL model was 12 inches long with a sphericallyblunted 8-degree cone forebody. The spherical nosewas 1 inch in diameter. The conical forebody inter-sected a cylindrical aft section which had cuts takenout of it at four circumferential locations 90 degrees toeach other. Flaps were located on each cut.

For this comparison, the LAURA CFD code wasused for obtaining computational predictions. LAURAis an upwind-biased, point-implicit, three-dimensional,Navier-Stokes algorithm.26 For the predictions pre-sented here, the code was run in a laminar, perfectgas, thin-layer Navier-Stokes mode.

The VT/VL model was tested in the 31-Inch Mach10 Tunnel at 0.0, 5.0, 17.5 and 25.0 degrees angleof attack. The data were non-dimensionalized withthe stagnation point heating to a 1-inch diameterhemisphere, as calculated with the Fay-Riddell theory.CFD predictions were calculated using tunnel run con-ditions at each angle angle of attack. The CFD resultsat 17.5� angle of attack were run through the RE-GRID program previously described and the resultsare shown in Figure 15. In the �gure, good agreementon the conical forebody can be observed although backon the ap (de ected at 20�), the phosphor data showssubstantially higher heating. The disagreement on the ap is believed to be due to the separated free-shearlayer from the cone transitioning and reattaching onthe ap, a ow situation which is challenging for CFDcomputations to predict.

Centerline cuts were also obtained at each angle ofattack on the conical forebody only and the resultscomparing the CFD predictions with the phosphordata are shown in Figure 16. The results at the higherangles of attack show an initial drop in heating fol-lowed by a slight rise due to the occurance of anover-expansion followed by a recompression in the noseregion of the model. In general, at all angles of attack,

14 of 20

American Institute of Aeronautics and Astronautics Paper 98{0712

Page 16: paper98...AIAA 98–0712 Reduction and Analysis of Phosphor Thermography Data With the IHEAT Software Package N. Ronald Merski NASA Langley Research Center, Hampton, VA 23681 36th

0 2 4 6 8x (inches from nose leading edge)

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8h/

h FR

Phosphors α=25.0°LAURA α=25.0°Phosphors α=17.5°LAURA α=17.5°Phosphors α=5.0°LAURA α=5.0°Phosphors α=0.0°LAURA α=0.0°

Fig. 16 VT/VL Mach 10 windward centerline

heating comparison between phosphor data and

LAURA computations.

Centerlinex/L=.20

x/L=1.0

x/L=.78

Fig. 17 X-34 planform with pro�le cut locations.

there is very good agreement between the phosphordata and CFD predictions.

Mach 6 X-34 Comparison

CFD comparisons have also been made on an X-34con�guration (planform shown in Figure 17) at Mach6 conditions. For this study, 0.0183-scale phosphormodels were fabricated and tested in the 20-Inch Mach6 Tunnel at freestreamReynolds numbers varying from0.59 to 7.95�106=ft: as described in Ref. 4.

The CFD computations were performed usingLAURA, LATCH and GASP as described in Refs. 27,28, and 4 respectively. The codes were run in laminarand turbulent perfect gas modes at tunnel conditionscorresponding to a freestream Reynolds number of7.86�106=ft. The LAURA and the GASP calculationsused the same grid which was shock-adapted from theLAURA code.

REGRID comparisons between the laminar LAURAand GASP results and phosphor data are shown inFigs. 18 and 19. Turbulence results are shown inFigs. 20 and 21.

In the images, the phosphor data is seen to have asharp jump in heating due to the onset of transition

0 0.125 0.500

h/hFR

0.3750.250

LAURA Prediction

Phosphor Data

Fig. 18 Laminar Phosphor/LAURA X-34 compar-

ison, � = 15�.

0 0.125 0.500

h/hFR

0.3750.250

GASP Prediction

Phosphor Data

Fig. 19 Laminar Phosphor/GASP X-34 compari-

son, � = 15�.

0 0.125 0.500

h/hFR

0.3750.250

LAURA Prediction

Phosphor Data

Fig. 20 Turbulent Phosphor/LAURA X-34 com-

parison, � = 15�.

15 of 20

American Institute of Aeronautics and Astronautics Paper 98{0712

Page 17: paper98...AIAA 98–0712 Reduction and Analysis of Phosphor Thermography Data With the IHEAT Software Package N. Ronald Merski NASA Langley Research Center, Hampton, VA 23681 36th

0 0.125 0.500

h/hFR

0.3750.250

GASP Prediction

Phosphor Data

Fig. 21 Turbulent Phosphor/GASP X-34 compar-

ison, � = 15�.

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

x/L

0

0.05

0.1

0.15

0.2

0.25

0.3

h/h F

R

Phosphor DataLAURA (Laminar)LAURA (Turbulent)GASP (Laminar)GASP (Turbulent)LATCH (Laminar)LATCH (Turbulent)

Fig. 22 Phosphor/CFD X-34 centerline compari-

son.

starting a third of the way back along the centerline.In comparing the phosphor data to the computationalresults, the laminar predictions should only be com-pared with the laminar phosphor data and the sameis true for the turbulent predictions. The images showthat the GASP predictions appear to run higher thanthe LAURA predictions in both the laminar and tur-bulent cases.Line cuts were taken from the phosphor and CFD re-

sults along the centerline, and at axial stations (shownin Fig. 17) of x/L=0.2 and 0.78, where L is de�nedto be from the nose to the junction between the apand the fuselage. The centerline phosphor heating inFig. 22 in the nose is laminar. A minimum is seenin the heating and then the ow transitions to tur-bulence from an axial station of 0.25 to about 0.55.Agreement between the LAURA predictions and thephosphor data in both the laminar and turbulent re-gions is within experimental uncertainty. The GASP

0 0.1 0.2

2y/b

0

0.05

0.1

0.15

0.2

h/h F

R

Phosphor DataLAURAGASPLATCH

Fig. 23 Laminar phosphor/CFD X-34 comparison

at x/L=0.02.

data tends to agree in the laminar region, but is higherthan the phosphor results in the turbulent region.The LATCH results split the di�erence between theLAURA and GASP results in the turbulent region butare higher than the phosphor data and LAURA andGASP predictions in the laminar region.

Figure 23 shows laminar comparisons at the 0.2 axialstation with respect to 2y=b where y is the spanwiselocation and b is the total span from wing tip to wingtip. The LAURA predictions agree with the phosphordata to within the data uncertainty for most of the cutwhile the GASP predictions tend to run a little higherthan the phosphor data and the other predictions. TheLATCH predictions start o� higher than the phosphordata near the centerline and end up low outboard onthe fuselage.

In the case of the 78 percent cut, Fig. 24, accuratelypredicting the heating is more di�cult because of thepresence of the bow-shock/wing-shock interaction onthe wing. The LAURA predictions are in qualitativeagreement with the phosphor results. Along the cutfrom the centerline, the heating experiences a rise asthe heating from the wing fuselage junction is reached.Part of the footprint of the shock impingement is seenby the the two lower humps in the phosphor data andthe single lower hump in the LAURA prediction. Fi-nally, a rise is seen toward the wing tip in both thephosphor and LAURA results with a slight drop inthe phosphor results due to the fact that a laminarportion of the wing is being cut. Qualitatively andquantitatively, the results are not quite as good withthe GASP and LATCH results. Both of these codestended to smear out the characteristics associated withthe shock interaction.

16 of 20

American Institute of Aeronautics and Astronautics Paper 98{0712

Page 18: paper98...AIAA 98–0712 Reduction and Analysis of Phosphor Thermography Data With the IHEAT Software Package N. Ronald Merski NASA Langley Research Center, Hampton, VA 23681 36th

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

2y/b

0

0.1

0.2

0.3

0.4h/

h FR

Phosphor DataLAURAGASPLATCH

Fig. 24 Turbulent phosphor/CFD X-34 compari-

son at x/L=0.78.

Application of Extrapolation Analysis

to the X-34

Based on the relatively good agreement betweenthe phosphor data and the LAURA computations attunnel conditions, the extrapolation theory previouslydescribed was applied to the X-34 phosphor data andcompared to LAURA ight surface temperature pre-dictions. Phosphor data at � = 23� were used, sinceboth laminar and turbulent predictions were availableat this angle of attack.

The phosphor data was extrapolated to ightsurface temperatures using the EXTRAP program.Flight conditions for the comparison were at an al-titude of 118,419ft: and a velocity of 6490ft=s whichcorresponded to Mach 6.4 and a length Reynolds num-ber of 16.2�106. In the case of the laminar data, a runwas used with Re1 = 1:0�106=ft:With the turbulentextrapolation, a Reynolds number de�cit existed be-tween the tunnel conditions and the ight conditions,so two runs were used at Re1 = 4:4 � 106=ft: and7:8� 106=ft: to show the applicability of the methodto di�erent tunnel Reynolds numbers.

Laminar Extrapolation

The global comparison between the laminar phos-phor data and the LAURA computations is shown inFig. 25 along with a centerline cut and an axial cut atx/L=0.78 in Figs. 26 and 27 respectively.

Generally the surface temperature levels comparewell over the whole image. On the wing, the pat-terns due to the shock impingement appear somewhatsharper with the phosphor data than with the CFDpredictions, possibly due to lack of grid resolution onthe part of the CFD computatons. On the centerlinecut, the CFD curve is atter than the phosphor curvewith the phosphor results meandering on either side

0 562 2250

Tw, °F

11681125

LAURA Prediction

Phosphor Data

Fig. 25 Laminar surface temperatures at � = 23�.

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

x/L

0

500

1000

1500

2000

Tem

per

atu

re(°

F)

LAURA PredictionPhosphor Data

Fig. 26 Laminar centerline ight surface temper-

atures.

of the CFD predictions, the worst disagreement beingabout 100�F towards the aft end of the con�guration.In the case of the axial cut, good agreement is ob-served with the exception of the fact that the smearedshock footprint in the CFD computation lowers thepeak temperature predicted on the wing by approxi-mately 150�.

Turbulent Extrapolation

The global comparison between the turbulent phos-phor data extrapolated from the 8�106=ft: freestreamReynolds number case and the LAURA computationsis shown in Fig. 28. The rearward part of the fuselagein the compuational results is not shown since it wasnot available at the time the extrapolations were per-formed. In the turbulent case, good agreement in thesurface temperature levels is generally observed withthe exception of the region outboard of the interactionregion and perhaps near the wing leading edge wherethe phosphor data predicts slightly higher tempera-

17 of 20

American Institute of Aeronautics and Astronautics Paper 98{0712

Page 19: paper98...AIAA 98–0712 Reduction and Analysis of Phosphor Thermography Data With the IHEAT Software Package N. Ronald Merski NASA Langley Research Center, Hampton, VA 23681 36th

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

2y/b

0

500

1000

1500

2000T

emp

erat

ure

(°F

)

LAURA PredictionPhosphor Data

Fig. 27 Laminar ight surface temperatures at

x/L=0.78.

0 562 2250

Tw, °F

11681125

LAURA Prediction

Phosphor Data

Fig. 28 Turbulent surface temperatures at � =

23�.

tures. Line cuts are shown on the centerline and at anaxial station of x/L=0.78 in Figs. 29 and 30 respec-tively. For the phosphor data, extrapolated resultsare shown from two runs, at Re1 = 4:4� 106=ft: and7:8� 106=ft: Agreement is seen to within 75�F every-where in the two plots where laminar ow exits on thewind tunnel model.

As previously stated, the purpose in using extrapo-lated data from tunnel runs at two di�erent freestreamReynolds numbers was to show the versatility of theextrapolation theory, in spite of a Reynolds numberde�cit between the tunnel and the ight conditions.Figure 31, which shows centerline cuts of unextrapo-lated wind tunnel data at Re1 = 4:4 � 106=ft: and7:8� 106=ft:, is presented to illustrate the fact thereis a signi�cant di�erence (as much as 30 percent) inthe heat transfer levels in the turbulent region be-tween the two Reynolds number conditions. Without

0 0.25 0.5 0.75 1

x/L

0

250

500

750

1000

1250

1500

1750

2000

2250

2500

Tem

pera

ture

(°F

)

LAURA PredictionExtrapolation, tunnel Re∞=7.8x106/ft.Extrapolation, tunnel Re∞=4.4x106/ft.

Fig. 29 Turbulent centerline ight surface tem-

peratures at � = 23�.

0 0.25 0.5 0.75 1

x/L

0

250

500

750

1000

1250

1500

1750

2000

Tem

pera

ture

(°F

)

LAURA PredictionExtrapolation, tunnel Re∞=7.8x106/ft.Extrapolation, tunnel Re∞=4.4x106/ft.

Fig. 30 Turbulent ight surface temperatures at

x/L=0.78.

the turbulent extrapolation factor previously derived,this would correspond to over a 120�F di�erence inpredicted ight surface temperature levels. This dif-ference is not observed on the centerline cut betweenthe two Reynolds number runs.

Concluding Remarks

This paper has presented the basic process whichis used at the Langley Research Center in obtainingphosphor thermography data in hypersonic facilities.A new weighted relative-intensity uorescence theoryallows for the quantitative determination of surfacetemperature measurements on complex models havinglarge of suface curvature. Application of an enthalpy-based approach to solving the heat conduction equa-tions along with calibration of the e�ective starting

18 of 20

American Institute of Aeronautics and Astronautics Paper 98{0712

Page 20: paper98...AIAA 98–0712 Reduction and Analysis of Phosphor Thermography Data With the IHEAT Software Package N. Ronald Merski NASA Langley Research Center, Hampton, VA 23681 36th

0.25 0.5 0.75

x/L

0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45h/

h FR

Re∞=7.8x106/ft.Re∞=4.4x106/ft.

Fig. 31 Tunnel centerline heat transfer coe�cients

at Re1 =4.4 and 7.8x106/ft.

time of a run and a simple approach for account-ing for variable substrate thermal properties improvesthe accuracy of global heat transfer computations. Inaddition, a method has been developed for extrapolat-ing wind tunnel data to ight surface temperatures.A user-friendly set of GUI-driven codes, collectivelyknown as IHEAT, allows for the rapid reduction andanalysis of the large amount of image data associ-ated with phosphor thermography. Coupled with thisprocess is a rapid ceramic model casting technique.The result is a methodology which quickly provides awealth of information critical to the design of a ther-mal protection system including e�ects of Reynoldsnumber, Mach number, angle of attack, control sur-face de ection and sideslip angle.

Acknowledgements

The following people are gratefully acknowledged fortheir contributions to this work: Chris Riley for hisLAURA VT/VL centlerline solutions; Scott Berry andTom Horvath for the X-34 phosphor data; Bill Woodand Chris Glass for the 15-degree laminar and turbu-lent X-34 LAURA and GASP solutions, respectively;and Bill Kleb for the laminar and turbulent 23-degree ight solutions.

References1Berry, S. A., Bouslog, S. A., Brauckmann, G. J., and

Caram, J. M., \Boundary Layer Transition Due to Isolated

Roughness: Shuttle Results From the LaRC 20-Inch Mach 6

Tunnel," AIAA Paper 97{0273, Jan. 1997.2Thompson, R., Hamilton, H., Berry, S., T.J.Horvath, and

Nowak, R., \Hypersonic Boundary-Layer Transition for X-33

Phase II Vehicle," AIAA Paper 98{0867, Jan. 1998.3Hamilton, H. H., Weilmuenster, K. J., Horvath, T. J., and

Berry, S. A., \Computational/Experimental Aeroheating Pre-

dictions for X-33 Phase II Vehicle," AIAA Paper 98{0869, Jan.

1998.

4Berry, S. A., Horvath, T. J., DiFulvio, M., Glass, C. E.,

and Merski, N. R., \X-34 Experimental Aeroheating at Mach 6

and 10," AIAA Paper 98{0881, Jan. 1998.5Campbell, C. H., Caram, J. M., Berry, S. A., Horvath,

T. J., and Merski, N. R., \An Overview of X-38 Hypersonic

Aerothermodynamic Wind Tunnel Data and Comparison with

Numerical Results," AIAA Paper 97{2475, Jun. 1997.6Berry, S. A., Horvath, T. J., and Williams, G. B., \Results

of Aerothermodynamic and Boundary Layer Transition Testing

of 0.0362-Scale X-38 (Rev 3.1) Vehicle in NASA Langley 20-Inch

Mach 6 Tunnel," NASA TM 112857, Sep. 1997.7Loomis, M. P., Venkatapathy, E., Papadopoulos, P., Davis,

C. B., Campbell, C., Berry, S. A., Horvath, T. J., and Merski,

N. R., \Aerothermal CFD Validation and Prediction for the X-

38 Program," AIAA Paper 97{2484, Jun. 1997.8Buck, G. M., \An Imaging System for Quantitative Surface

Temperature Mapping Using Two-Color Thermographic Phos-

phors," ISA Paper 88{0772, May 1988.9Buck, G. M., \Surface Temperature/Heat Transfer Mea-

surement Using a Quantitative Phosphor Thermography Sys-

tem," AIAA Paper 91{0064, Jan. 1991.10Merski, N. R., \A Relative-Intensity Two-Color Phosphor

Thermography System," NASA TM 104123, Sep. 1991.11Merski, N. R., \An Improved Two-Color Relative-Intensity

Phosphor Thermography Method For Hypersonic Wind Tunnel

Aeroheating Measurements," Dsc dissertation, 1998.12Garlick, G., Luminescent Materials, Clarendon Press, Ox-

ford, 1949.13Thomas, A. C. and Perlbachs, A. P., \Application of

Ground Test Data to Reentry Vehicle Design," Tech. Rep.

AFFDL-TR-66-229, Air Force Flight Dynamics Laboratory,

Jan. 1967.14Tauber, M. E. and Meneses, G. P., \Aerothermodynamics

of Transatmospheric Vehicles," AIAA Paper 86{1257, Jun. 1986.15Anderson, J. D., Hypersonic and High Temperature Gas

Dynamics, McGraw-Hill Book Company, 1989.16McBride, B. J., Gordon, S., and Reno, M. A., \Coe�cients

for Calculating Thermodynamic and Transport Properties of In-

dividual Species," NASA TM 4513, Oct. 1993.17U.S. Standard Atmosphere, National Oceanic and Atmo-

spheric Administration, NASA and U.S. Air Force, Washington,

DC, 1976.18Buck, G. M. and Vasquez, P., \An Investment Ceramic

Slip-Casting Technique for Net-Form, Precision, Detailed Cast-

ing of Ceramic Models," U.S. Patent 5,266,252, Nov. 1989.19Micol, J. R., \Langley Aerothermodynamic Facilities Com-

plex: Enhancements and Testing Capabilities," AIAA Paper

98{0147, Jan. 1998.20AIAA Standard: Assessment of Wind Tunnel Data Un-

certainty, American Institute of Aeronautics and Astronautics,

Washington, D.C., AIAA Standard No. S-071-1995, 1995.21Guide to the Expression of Uncertainty in Measurement,

International Organization for Standardization, Switzerland,

Corrected Edition, ISBN 92-67-10188-9, 1995.22Hollis, B. R., \User's Manual for the One-Dimensional Hy-

personic Experimental Aero-Thermodynamic (1DHEAT) Data

Reduction Code," NASA CR 4691, Aug. 1995.23Kendall, D. N., Dixon, W. P., and Schulte, E. H.,

\Semiconductor Surface Thermocouples for Determining Heat-

Transfer Rates," IEEE Transactions on Aerospace and Elec-

tronic Systems, Vol. 3, No. 4, Jul. 1967, pp. 596{603.24Fay, J. A. and Riddell, F. R., \Theory of Stagnation Point

Heat Transfer in Dissociated Air," Journal of Aeronautical Sci-

ences, Feb. 1958, pp. 73{85.25Hamilton, H. H., Greene, F. A., and DeJarnette, F. R., \An

Approximate Method for Calculating Heating Rates on Three-

Dimensional Vehicles," AIAA Paper 93{2881, Jul. 1993.

19 of 20

American Institute of Aeronautics and Astronautics Paper 98{0712

Page 21: paper98...AIAA 98–0712 Reduction and Analysis of Phosphor Thermography Data With the IHEAT Software Package N. Ronald Merski NASA Langley Research Center, Hampton, VA 23681 36th

26Gno�o, P. A., \An Upwind-Biased, Point-Implicit Relax-

ation Algorithm for Viscous, Compressible Perfect-Gas Flows,"

Tech. Rep. TP-2953, NASA, Feb. 1990.27Kleb, W. L., Wood, W. A., Gno�o, P. A., and Alter, S. J.,

\Computational Aeroheating Predictions for X-34," AIAA Pa-

per 98{00879, Jan. 1998.28Riley, C. J., Kleb, W. L., and Alter, S. J., \Aeroheat-

ing Predictions for X-34 Using an Inviscid-Boundary Layer

Method," AIAA Paper 98{00880, Jan. 1998.

20 of 20

American Institute of Aeronautics and Astronautics Paper 98{0712