LIBRARY Michigan State Unlvarllty PLACE ll RETURN BOX to romovo this checkout from your mood. TO AVOID FINES Mum on or baton date duo. DA'I; DUE DATE DUE DATE DUE ll NEW-r69 ] : l ———l A; :1 1 l MSU I. An Affirmative Action/Equal Opponunuy Institution Wanna-9.1 MULTI-DIMENSIONAL ESTIMATION OF THERMAL PROPERTIES AND SURFACE HEAT FLUX USING EXPERIMENTAL DATA AND A SEQUENTIAL GRADIENT METHOD By Kevin J Dowding A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Mechanical Engineering 1997 ABSTRACT MULTI-DIMENSIONAL ESTIMATION OF THERMAL PROPERTIES AND SURFACE HEAT FLUX USING EXPERIMENTAL DATA AND A SEQUENTIAL GRADIENT METHOD By Kevin J Dowding Inverse theory is an emerging field of study with application to a diverse range of problems. Inverse thermal problems are the focus of this dissertation; specifically, the parameter estimation problem and inverse heat conduction problem (H-ICP) are investi- gated. Although one-dimensional inverse thermal problems have been widely investigated, multi-dimensional problems are beginning to receive an increasing amount of attention. One- and two-dimensional cases are addressed for both the noted inverse problems, including an experimental application. Parameter estimation techniques are applied to estimate the thermal properties of a carbon-carbon composite from transient experiments. Properties are determined as a func- tion of temperature and direction relative to the fiber orientation. The thermal conductivity is assumed to be orthotropic, varying in the direction normal and parallel to the fibers; the volumetric heat capacity is assumed isotropic. Thermal properties from room temperature up to 500C are obtained. The thermal conductivity normal to the fiber is found to be less than one-tenth of the thermal conductivity parallel to the fiber. Agreement within 7% is demonstrate between independent one- and two-dimensional results. A sequential-in-time implementation is proposed for a conjugate gradient method, utilizing an adjoint equation approach, to solve the IHCP. Because the IHCP is generally ill-posed, Tikhonov regularization is included to stabilize the solution. The proposed sequential method benefits from the efficiency and on-line capabilities of a sequential implementation, without requiring a priori information about the (unknown) surface heat flux. Aspects of the sequential gradient method are discussed and examined. Several promising features of the sequential gradient method are noted. Simulated one- and two- dimensional test cases are presented to study the sequential implementation. Numerical solutions are obtained using a finite difference procedure. Results indicate the sequential implementation has accuracy comparable to a standard whole domain solution, but in cer- tain cases requires significantly more computational time. Methods to improve the compu- tational requirements, which make the method competitive, are presented. to my wife iv ACKNOWLEDGMENTS The input and help of the my Ph.D. guidance committee, James Beck, Patti Lamm, Alex Diaz, and Craig Somerton, is acknowledged. Special thanks are due Patti Lamm for her patience working with an engineer, help with understanding the gradient methods, and availability to discuss the research. To my advisor James Beck, I thank for his encourage- ment and introduction to the field of inverse problems. His enthusiasm and dedication were inspiring, not only in helping realize this work, but as a teacher and researcher. I have benefited from interaction with several fellow students and researchers. Input and help from Arafa Osman, Bob McMasters, Heidi Relyea, and Matt White is appreci- ated. Completing this work would not have been possible if it were not for the endless sup- port my parents. To my wife I’m indebted for her understanding, support, and patience. TABLE OF CONTENTS LIST OF TABLES ix LIST OF FIGURES Y NOMENCLATURE xiv Greek ................................................................................................................................ xvi Subscripts ........................................................................................................................ xvii Superscripts .................................................................................................................... xviii Chapter 1 INTRODUCTION 1 Chapter 2 CARBON -CARBON THERMAL PROPERTIES: ONE-DIMENSIONAL EXPERI- MENTS 6 2-1.0 Introduction .............................................................................................................. 6 2-l.l Problem Description ............................................................................................................ 8 2-1.2 Literature review .................................................................................................................. 9 2-2.0 Experimental Aspects ............................................................................................ 11 2-2.1 Experimental Description .................................................................................................. 11 2-2.2 Experimental Design .......................................................................................................... 19 2-3.0 Analysis Procedures ............................................................................................... 21 2-4.0 Results and Discussion .......................................................................................... 23 2-4.1 Effective Properties for Mica Heater and Insulation .......................................................... 24 2-4.2 Effective Properties of Carbon-Carbon .............................................................................. 27 2-5.0 Experimental Uncertainty ...................................................................................... 36 Chapter 3 CARBON-CARBON THERMAL PROPERTIES: TWO-DIMENSIONAL EXPERI- MENTS 38 3-1.0 Introduction ............................................................................................................ 38 3-1.1 Motivation .......................................................................................................................... 38 3-2.0 Experimental Aspects ............................................................................................ 40 vi 3-3.0 Analysis Procedures ............................................................................................... 43 3-4.0 Results and Discussion .......................................................................................... 44 3-4.1 Experimental Data ............................................................................................................. 45 3-4.2 Estimated Thermal Properties ............................................................................................ 47 3-4.3 Comparison of One- and Two-Dimensional Results ......................................................... 60 3-5.0 Experimental Uncertainty ...................................................................................... 63 Chapter 4 SOLUTION OF IHCP USING A GRADIENT METHOD WITH ADJOINT EQUA- TION APPROACH 65 4-1.0 Introduction ............................................................................................................ 65 4-2.0 Problem Statement of the Multi-Dimensional IHCP ............................................. 70 4-3.0 Gradient Calculation .............................................................................................. 73 4-4.0 Sensitivity Equations ............................................................................................. 75 4-5.0 Adjoint Equations .................................................................................................. 77 4-6.0 Optimization Method ............................................................................................. 84 4-7.0 Implementation of the Gradient Method ................................................................ 87 4-7.1 Whole Domain ................................................................................................................... 87 4-7.2 Sequential Implementation ................................................................................................ 91 4-8.0 Parameterizing the Solution (Finite-Dimensional Problem) .................................. 96 4-9.0 Summary ................................................................................................................ 99 Chapter 5 APPLICATION OF THE SEQUENTIAL GRADIENT METHOD: ONE-DIMEN- SIONAL IHCP 101 5-1.0 Introduction .......................................................................................................... 101 5-2.0 Advantages of a Sequential Method and Numerical Solution of the IHCP ......... 103 52.] Numerical Solution .......................................................................................................... 103 5-2.2 Sequential Solution of IHCP ............................................................................................ 104 5-3.0 Simulated Test Cases ........................................................................................... 107 5-4.0 1D Results - Simulated Measurements ................................................................ 115 5-4.1 Exact Data (no Measurement Errors) .............................................................................. l 15 5-4.2 Corrupted Data (Measurement Errors) ............................................................................ 126 5-4.3 Results - Experimental Measurements ............................................................................. 140 Chapter 6 APPLICATION OF THE SEQUENTIAL GRADIENT METHOD: TWO-DIMEN- SIONAL IHCP 145 vii 6-1.0 Introduction .......................................................................................................... 145 6-2.0 Simulated Temperature Data ................................................................................ 146 6-3.0 2D Results - Simulated Measurements ................................................................ 149 6-3.1 Exact Data (No Measurement Errors) .............................................................................. 149 6-3.2 Corrupted Data (Measurement Errors) ............................................................................ 159 6-4.0 Results - Experimental Measurements ................................................................. 175 Chapter 7 SUMMARY AND CONCLUSIONS 185 Chapter 8 RECOMMENDED FUTURE WORK 187 APPENDIX A FINITE CONTROL VOLUME METHOD 189 A-1.0 Introduction .......................................................................................................... 189 A-2.0 Alternating Direction Implicit Method (ADI) ..................................................... 195 A-2.l ADI Equations (n+l/2) .................................................................................................... 196 A-2.2 ADI Equations (n+1) ....................................................................................................... 198 A-2.3 Summary of ADI Equations ............................................................................................. 200 LIST OF REFERENCE 201 viii Table 2-1 Table 2-2 Table 2-3 Table 2-4 Table 2-5 Table 3-1 Table 3—2 Table 5-1 Table 5-2 Table 5-3 Table 5-4 Table 6-1 Table 6-2 Table 6-3 Table 6-4 LIST OF TABLES Sensor locations in experimental apparatus ............................................... 14 Thermal properties of the ceramic insulation ............................................ 25 Effective thermal properties for the mica heater assembly ........................ 25 Thermal properties estimated for the carbon-carbon composite from one- dimensional experiments and analysis ....................................................... 27 Experimental uncertainty for the two-dimensional experiments ............... 37 Thermal properties estimated for the carbon-carbon composite from two- dimensional experiments and analysis ....................................................... 48 Experimental uncertainty for the two-dimensional experiments ............... 63 Estimation results for exact simulated data ............................................. 117 Estimation results for simulated data corrupted with errors .................... 130 Estimation results for prescribing a constant heat flux over the sequential interval ..................................................................................................... 138 Estimation results for analysis of experimental case ............................... 144 Estimation results for the two-dimensional IHCP with exact simulated data ........................................................................................................... 155 Estimation results for two-dimensional IHCP with simulated data corrupted with random errors ................................................................... 162 Estimation results for two-dimensional IHCP with simulated data corrupted with random errors using prior information ............................ 174 Estimation results for two-dimensional IHCP with experimentally measured heat flux ................................................................................... 183 ix Figure 2-1 Figure 2-2 Figure 2-3 Figure 2-4 Figure 2-5 Figure 2-6 Figure 2-7 Figure 3-1 Figure 3-2 Figure 3-3 Figure 3-4 Figure 3-5a Figure 3-5b Figure 3-5c Figure 3-6a LIST OF FIGURES Schematic of experimental set-up used for estimating thermal properties of carbon-carbon composite. For one-dimensional experiments all heaters are activated. .............................................................................................. 12 Heat transfer model for one-dimensional experiment ............................... 15 Temperature dependence of thermal properties estimated from one- dimensional experiments ........................................................................... 29 Experimental data for one-dimensional case 1023#508.2 ......................... 30 Sequential parameter estimates for one-dimensional case 1023#508.2 ...32 Temperature residuals for one-dimensional case 1023#508.2 ................... 33 Sensitivity coefficients for one-dimensional case 1023#508.2 .................. 35 Heat transfer model for two-dimensional experiments .............................. 41 Experimental data for two-dimensional case 1022$297, see Table 2-1 for sensor locations .......................................................................................... 45 Temperature dependence of estimated thermal properties from two- dimensional experiments ........................................................................... 49 Sequential parameter estimates for two-dimensional case 1022$297 ....... 50 Temperature residuals for two-dimensional case 1022$297 on surface below active heater ..................................................................................... 52 Temperature residuals for two-dimensional case 1022$297 on the heated surface, but not on the are of the active heater ........................................... 53 Temperature residuals for two-dimensional case 1022$297 at the specimen/insulation interface ..................................................................... 54 Normalized sensitivity coefficient for volumetric heat capacity for two- dimensional case 1022$297 ....................................................................... 57 Figure 3-6b Figure 3-6c Figure 3-7 Figure 4-1 Figure 5-1 Figure 5-2 Figure 5-3 Figure 5-4 Figure 5-5a Figure 5-5b Figure 5-5c Figure 5-6 Figure 5-7 Figure 5-8 Figure 5-9 Figure 5-10 Normalized sensitivity coefficient for thermal conductivity in y-direction for two-dimensional case 102233297 .......................................................... 58 Normalized sensitivity coefficient for thermal conductivity in x-direction for two-dimensional case 1022$297 .......................................................... 59 Comparison of thermal properties estimated from one— and two- dimensional experiments a) thermal conductivity and b) volumetric specific heat ................................................................................................ 62 Schematic of multi-dimensional general IHCP ......................................... 70 Simulated temperature and heat flux data for the one-dimensional triangular heat flux ................................................................................... 112 Simulated temperature and heat flux data for the one-dimensional step heat flux test case ..................................................................................... 1 13 Simulated temperature and heat flux data for the one-dimensional sinusoidal heat flux test case .................................................................... 1 14 Estimated heat flux for one-dimensional triangular test case with exact data ........................................................................................................... 116 Estimated one-dimensional triangular heat flux with no measurement errors for sequential interval beginning at time 0.12 for r = 5 ................. 123 Estimated one-dimensional triangular heat flux with no measurement errors for sequential interval beginning at time 0.12 for r = 8 ................. 124 Estimated one-dimensional triangular heat flux with no measurement errors for sequential interval beginning at time 0.12 for r = 10 ............... 125 Estimated heat flux for one-dimensional triangular test case with measurement errors .................................................................................. 127 Estimated heat flux for one-dimensional step test case with measurement errors ........................................................................................................ 128 Estimated heat flux for one-dimensional sinusoidal test case with measurement errors .................................................................................. 129 Procedure for specifying the initial condition using converged estimates from the previous sequential interval ....................................................... 135 Estimated heat flux for one-dimensional triangular case with the heat flux held constant over future sequential interval ........................................... 139 xi Figure 5-11 Figure 6-1 Figure 6-2 Figure 6—3 Figure 6-4a Figure 6-4b Figure 6-5a Figure 6-5b Figure 6-6a Figure 6—6b Figure 6-7 Figure 6-8 Figure 6—9a Figure 6-9b Figure 6-9c Figure 6-9d Estimate heat flux for one-dimensional experimental case 1010#30.1 ..143 Two-dimensional geometry for simulated test cases ............................... 147 Simulated temperature and surface heat flux for a two-dimensional triangular heat flux ................................................................................... 150 Simulated temperature and surface heat flux for a two-dimensional step heat flux .................................................................................................... 151 Estimated surface two-dimensional heat flux for triangular test case using data without measurement errors. Whole domain solution ..................... 153 Estimated surface two-dimensional heat flux for triangular test case using data without measurement errors. Sequential solution with 1:6 ............. 154 Estimated surface heat flux for triangular test case using data corrupted with measurement errors (0’ = 0.0018°C ). Whole domain solution. 160 Estimated surface heat flux for triangular test case using data corrupted with measurement errors (6 = 0.0018°C ). Sequential solution r=6 161 Estimated surface heat flux for step test case using data corrupted with measurement errors (0’ = 0.0025°C ). Whole domain solution ............. 164 Estimated surface heat flux for step test case using data corrupted with measurement errors (0' = 0.0025°C ). Sequential solution r=15 ........... 165 Estimated surface heat flux for triangular test case using data corrupted with measurement errors ( 0' = 0.0018°C ). Sequential solution r=15 168 Estimated surface heat flux for triangular test case using data corrupted with measurement errors (6 = 0.0018°C ). Sequential solution with $6 using prior information. ........................................................................... 173 Measured surface heat for experimental case to estimate thermal properties of carbon-carbon ..................................................................... 177 Estimated surface heat flux for experimental case. Whole domain ......... 180 Estimated surface heat flux for experimental case. Sequential solution with r-—10. ................................................................................................. 181 Estimated surface heat flux for experimental case. Combined function specification regularization method with r=6. ......................................... 182 xii Figure A-l Computational nodes and control surfaces .............................................. 191 Figure A-2 Typical energy balance for an interior node ............................................. 191 Figure A-3 Finite control volumes along the boundary of the domain ...................... 194 xiii ”f." NOMENCLATURE dimension of two-dimensional simulated case [m] area [m2] sub-diagonal of tri-diagonal matrix dimension of two-dimensional simulated case [m] parameter vector estimated parameter vector sup-diagonal of tri-diagonal matrix specific heat [J/kgC] main diagonal of tri-diagonal matrix location for temperature sensor j [m] temperature residual sensor j [°C] function space nonhomogeneous term for boundary surface [1“,] convection coefficient [W/(m2°C)] boundary coefficient number of components retained in sequential implementation current [Amps] objective function [°C2 sec] sum-of-squares term in objective function [°Czsec] regularization (Tikhonov) term if objective function [°Czsec] xiv *0 23: => ”6 a qprr'or qest thermal conductivity [W/m° C] boundary coefficient length [m] function space of all “square integrable” functions number of temporal components for estimated heat flux on [F4] outward pointing unit normal vector number of measurement times number of estimated parameters number of parameters search direction [—o—:——:| (W/m )m number of spatial components for heat flux on [F4] heat flux [W/mz] estimated heat flux [W/mz] estimated heat flux with measurement errors [W/mz] prior information for heat flux, equation (4—2.4) [W/mz] stored flux for sequential solution [W/mz] two-dimensional coordinate vector [m] number of future time steps sum-of-squares function [°C2] temperature error, equation (5-4.2) [°C] basis for finite-dimensional heat flux, equation (4-7.6) time [sec] heating time [sec] temperature [°C] XV ?<€< >< ?O (2'2) 8T . 42...?!“ = qm (2-3., y y = 0 Tmica(y,0) = To, —LmicaSySO (2-3b) The second problem is for the carbon-carbon composite ke 66— Ce 8T“ 0 L t>O 24 y,ch _(p )CC-g; ’ 0 (2-6) aTins _a_, = 0 (2-7a) } r=Lm+A Tim.(y, 0) = To, L), S y _<. Ly + Li,” (2-7b) The problems are coupled through interface conditions that assume perfect contact ke arm,m(o,z)_ e BTCC(O,t) mica av _ .y, CC ay 9 (2'83) 17 Tmica(0’ t) = ch(0’ t) (2-8b) 6T," (L ,t) aT (L ,t) kins Say y = k; cc CCay)’ (2-8C) Tins(Ly’ t) = ch(Ly’ I) (2-8d) Although perfect contact is not reasonable between the mica heater and carbon-carbon, effective properties are used to represent the mica heater and contact conductance due to imperfect contact. Effective thermal conductivity is defined 8T ke : —/— - q a)? (2 9) _ . T . . where q rs the average heat flux and 9— 13 the average temperature gradient. Both aver- 3y ages are in the direction normal to the heat flow. The effective volumetric heat capacity is defined (pC)"’ = ‘l/IpCdv (2-10) v where V is the volume. By experimentally measuring effective properties for the mica heater assembly and insulation the contact conductance due to imperfect contact is accounted for in the model. In addition, the effective properties of the mica heater account for the effects of other materials in the heater and the cement used to install the thermo- couples in grooves in the carbon-carbon composite. Modeling the effects of several mate- rials with an effective property is much easier (and in most cases more accurate) then accounting for all effects individually. Effective properties of the carbon-carbon represent the non-homogenous construction of the material. To protect from oxidation a thin layer (approximately 10% of the overall thickness) of silicon-carbide protects the outer surfaces 18 of carbon-carbon. Consequently, effective properties that represent the carbon-carbon and silicon-carbide are estimated. Including the mica heater and ceramic insulation in the model (Figure 2-2) requires the thermal properties of the materials to be known, or also estimated, to deter- mine the properties of the carbon-carbon specimen. Neglecting the mica heater in the model is not appropriate because the contact conductance results in a large temperature drop across the heater. If the heater is neglected, the carbon-carbon properties will incor- rectly reflect this effect. Also, including the heat loss to the insulation, instead of assuming a perfect insulated condition (at the specimen/insulation interface), increases the accuracy of the properties estimated for the carbon-carbon. If known (tabulated or published) prop- erties are used for these materials (mica and insulation), several problems arise. First, ther- mal properties are typically not known very accurately. Second, contact conductance between adjacent layers is typically not negligible and must be considered. Third, the ceramic insulation was treated by spraying with a ridigizing material, possibly changing its thermal properties. By experimentally estimating effective properties of these materi- als, these problems are less influential in the estimated properties of the carbon-carbon. This approach requires additional experimental work, however. The experimental conditions (length of experiment and heating duration) to determine the properties of the mica and insulation are quite different from the conditions necessary to estimate the prop- erties of the carbon-carbon. A separate series of tests are performed, one set with the car- bon-carbon specimen removed, to determine the properties of the insulation and mica heater assembly. These are effective properties, which will account for any contact con- ductance due to imperfect contact between the different material layers. 19 With the separate series of experiments the effective thermal properties of the ceramic insulation and mica heater assembly were determined. However, when the set-up is recon- figured to test the carbon-carbon composite, the contact conductance may vary. Therefore, for each experiment to determine the properties of the carbon-carbon composite, a short duration experiment is conducted to characterize the effective thermal properties of the mica heater and contact conductance. The short duration experiment is conducted such that the thermal properties of the mica heater, including the contact resistance, are impor- tant in the thermal model, but the properties of the carbon-carbon are not important. Importance of the materials in the model is quantified by the sensitivity coefficients (dis- cussed in Section 23.0). A duration is selected that is sensitive to the properties of the mica, but relatively insensitive to the properties of the properties of the carbon-carbon. For a typical set of experiments to determine the properties of the carbon-carbon composite, the furnace is set at a given temperature and the experimental apparatus is allowed to reach a uniform temperature (usually this required several hours). A short dura- tion experiment (heating for approximately 2 seconds) is run first, which is used to esti- mate the effective properties of the mica heater. After allowing the set-up to stabilize at a uniform temperature, a second experiment is run. The second experiment (heating for approximately 20 seconds) is longer in duration and is used to estimate the properties of the carbon-carbon composite specimen. 2-2.2 Experimental Design During the initial stages of this research a different heater design was used. The heater was constructed of a Kapton material with resistance temperature sensors (RTD) integral in the heater assembly. This approach had the advantage of being non-intrusive and 20 required no machining to install thermocouples. A major difficultly with this design was that the magnitude of the contact conductance resulted in the one-dimensional thermal resistance of the heater/contact conductance being on the same order of magnitude as the composite specimen in the thermal model L L y z kapton (2_11) e e k J” CC kkapton where kzapton is the effective thermal conductivity of the kapton heater/contact conduc- tance and Lkapmn is its thickness. Hence, the measured RTD temperature depends as much on the thermal properties specified for the heater/contact conductance as it does on the thermal properties specified for the composite specimen. (Sensitivity coefficients, dis- cussed in Section 2-3.0, for the two materials would be comparable in magnitude.) In gen- era], when other materials are present in the model, their effect on the temperature should be small in comparison to the material for which the properties are sought. For this case, the sensitivity coefficients (discussed below) for k: C should be larger, compared to the c sensitivity coefficients for kzapwn . To improve the accuracy and reduce the dependence on the thermal properties of the heater/contact conductance, the sensors were embedded in the surface of the specimen. The choice of the boundary condition on the backside of the specimen is dependent on the thermal conductivity of the test specimen. For a test specimen with a relatively high thermal conductivity (which this carbon-carbon does have) it is more practical to simulate an insulation condition. An insulating boundary condition is less sensitive to the contact conductance at this surface. However, for a relatively low thermal conductivity material it is more practical to simulate a temperature boundary condition. For a test specimen with a 21 relatively low thermal conductivity, the boundary temperature is again less sensitive to the contact conductance. Based on optimality criteria for the one-dimensional case with an applied surface heat flux, a specified temperature boundary condition on the back surface is an optimal experiment for estimating thermal conductivity. An insulating condition is optimal for estimating the volumetric heat capacity, Beck and Arnold (1977). An approximate insulation boundary condition on the back of the specimen is chosen, instead of a temperature boundary condition (which was initially used), to minimize the sensitivity to the contact conductance at this location. With the insulation boundary condi- tion, the heat flux (and temperature gradients) at the specimen/insulation interface is small and the temperature is not sensitive to the contact conductance or location for the non- embedded sensors. A temperature boundary condition, however, has a larger heat flux at the boundary and the temperature (at the sensor) is very sensitive to the magnitude of the contact conductance and the specified location for the sensors at the interface. 2-3.0 Analysis Procedures The techniques to estimate thermal properties are detailed in a book by Beck and Arnold (1977). The basic process involves minimizing a sum-of-squares function 5 = (Y—T)TW(Y—T)+(u—b)TU(u—b) (2-12) where Y and T are vectors of the measured and calculated temperatures and W is a weighting matrix (typically the identity matrix). The last term in equation (2-12) serves as regularization or allows for'the inclusion of prior information about the thermal properties. It contains the difference between the prior information u and present estimates b with a symmetric weighting matrix U. To determine the thermal properties the function S is 22 minimized with respect to the thermal properties, i.e., 171 = (pC):c, b2 = k; cc, and b3 = ke This is accomplished by setting the first derivative of S with respect to each x,cc' parameter equal to zero, and solving for the estimated parameters (3). The resulting expression for the estimated thermal properties (Beck and Arnold, equation 7.4.6, 1977) is x(k+l) _ A k A A b b( )+ P(k)[XTW(Y — T) + U(u — b(k))] (2-13) P“) = [XTWX+ U1“ (2-14) The superscript (k) defines the iteration number, iteration is required even for the linear conduction problem, due to the non-linear nature of the estimation problem; the sensitivity coefficients X depend on the parameters (thermal properties). The columns of the sensi- tivity matrix are the first derivative of temperature (for each time and sensor location) with respect to the parameters X = [X19], sz, . . . ,Xbp] (2'15) .. 61 " ab. 1 X b (2-16) The sensitivity coefficients can provide considerable insight to the estimation problem and aid in the design of the experiment for optimum accuracy in the estimates (Beck and Arnold, Chapter 8, 1977). One criteria for an “optimal” experiment, valid for additive, zero mean normal errors in Y, and errorless independent variables, is to maximize A = IXTXI (2-17) This criteria is appropriate because it corresponds to minimizing the volume of the confidence region for the estimated parameters. By studying the experimental design prior to conducting experiments, such that equation (2-17) is maximized, the maximum information is available from an experiment. In addition to the criteria proposed in 23 equation (2-17), a constraint requiring a fixed number of observations and/or the same temperature rise, may be required to provide consistency in comparing experimental designs with multiple sensors. Taktak et al., (1993) discuss the design of an experiment to estimate thermal properties of composites with relatively low thermal conductivity. Additional investigations on experimental design concerning the optimal placement of sensors (Fadale et al., 1995a) and the design of experiments using uncertainty information (Fadale et al., 1995b and Emery and Fadale, 1996) use the Fisher information matrix. The sensitivity matrix and the temperatures are linearized about the thermal proper- ties from the previous iteration in equation (2-13) and (2-14). Iteration continues until con- kl Ak ak (+)_b() 3 8b“, vergence of the estimated parameters is reached, as defined by, |5 where e is a small number to quantify convergence, such as 8 = 0.0001 . These solution procedures are implemented with the computer program PROPlD to estimate the thermal properties. PROPID provides a means to estimate thermal proper- ties of multi-layer bodies from appropriate measurements. Thermal conductivity and volu- metric heat capacity may be determined simultaneously and for more than one material, if desired. Layers of different materials may also be lumped and effective thermal properties determined. Gamier et a1. (1992) performed experiments on materials with well-known published thermal properties to support the accuracy of PROPID. 2-4.0 Results and Discussion Three types of experiments were conducted. The first type removed the carbon-car- bon from apparatus and replaced it with a 1.25 cm thick insulation. These type-one exper- iments estimated effective thermal properties of the mica heater and insulation. 24 Experiments of type-two and type-three replaced the carbon-carbon specimen in the experiment. Type-two experiments heated for a short duration, providing information about the effective properties of the mica heater. Type-three experiments estimate the properties of the carbon-carbon. Results from experiments of type-one and type-two are discussed in Section 2-4.1 and type-three is discussed in Section 2-4.2. The reason that three types of experiments were used, is to quantify three separate, quite different, effects. Type-one experiments, with the carbon-carbon composite removed, provide effective thermal properties of the insulation. Experiment type-two was relatively short in duration, so that temperature is mainly influenced by the properties of mica heater, contact conductance, and associate materials used to install thermocouples. Type-three experiments were longer in duration and characterized the properties of the carbon-carbon composite. Each of the three experiments focuses on an individual aspect of the thermal model and has minimal effects from other material’s thermal properties. 2-4.1 Effective Properties for Mica Heater and Insulation An independent set of experiments was conducted to determine the effective thermal properties of the insulation material (Ulbrich, 1993). These tests used a set-up similar to Figure 2-1, except that the carbon-carbon specimen was replaced by a 1.25 cm-thick piece of insulation. The goal of these experiments was to estimate the thermal properties of the insulation. The results are given in Table 2-2. The values estimated for the insulation are considerably higher than the values reported by the manufacturer. This is not a surprising outcome. Manufacturers do not typically measure thermal properties. Also, this material was treated with a solution to make it more structurally strong. The effective properties for the mica heater were also measured. Because these values are sensitivity to the contact 25 Table 2-2 Thermal properties of the ceramic insulation Present Investigation Manufacturer Temperature kins (p C )fnsx 10-6 kins °C W/(m°C) J/(m3°C) W/(m°C) 40 .088 .419 175 .093 .570 200 .055 600 .1 10 Table 2-3 Effective thermal properties for the mica heater assembly Initial . kin-ca (pC)fm.wx10"6 Experimental Temp 0 3 Case (°C) (°C) W/(m°C) J/(m oC) 1010@3(),1 30 0.066 0.142 +/- 0.002 2.030 1010@41_1 41 0.070 0.131 +/- 0.002 2.030 1007@134,1 134 0.067 0.102 +/- 0.001 2.030 1007@143,1 143 0.072 0.1 10 +/- 0.002 2.030 1008@180,1 180 0.057 0.125 +/- 0.001 2.030 1012@195,1 195 0.053 0.123 +/- 0.001 2.030 1008@245.1 245 0.070 0.1 14 +/- 0.001 2.030 1013@254,1 254 0.057 0.123 +/- 0.002 2.030 1008@295,1 295 0.058 0.123 +/- 0.002 2.030 26 conductance they were measured again after the apparatus was re-assembled with the car- bon-carbon specimen. These experiments were referred to as type-one experiments. Having established the properties of the insulation, type-two experiments were con- ducted to determine the effective properties of the mica heater. It was found that estimat- ing the thermal conductivity and volumetric heat capacity simultaneously is not possible due to correlated sensitivity coefficients (see Beck and Arnold, 1977). Therefore, the volu- metric heat capacity of the mica heater is set at the value estimated during the experiments to determine the properties of the ceramic insulation (type-one). The effective thermal conductivity of the mica heater is estimated. The thermal conductivity is estimated because it is most influenced by the contact conductance, which may have changed from the previous estimates. The effective thermal conductivity estimated for the mica heater assembly, assuming (pC);,-ca = 2.0><106 J/m3C, demonstrated no specific trend with temperature over the range (30—295)°C . See Table 2-3. The largest value is ke (30°C) 2 0.14 W/mC mica e mica and the smallest is k (134°) = 0.10 W/mC . For temperatures up to 295°C the ther- mal conductivity of the mica heater is estimated at each temperature. After which, the magnitudes of these thermal properties were shown to minimally affect the outcome of the estimation for the carbon-carbon properties. A variation of 50% in the thermal properties specified for the mica heater results in variations in the estimated properties of the carbon- carbon that are within the magnitude of the confidence intervals (discussed below). For temperatures greater than 295°C , the thermal conductivity estimate of the mica heater at 295°C is used in the analysis. The fact that it is difficutlt to estimate the thermal proper- ties of the mica heater suggests that these properties are not significant in the estimation of 27 Table 2-4 Thermal properties estimated for the carbon-carbon composite from one-dimensional experiments and analysis . . e Experimental 113:2: 6 ky’ CC ( p C):c3x 10-6 Case (°C) (°C) W/(m°C) J/(m °C) 1010#30.1 31 0.128 3.40 +/- 0.05 1.42 +/- 0.01 1010#41.1 42 0.114 3.48 +/- 0.04 1.47 +/- 0.01 1012#94.1 95 0.161 3.85 +/— 0.07 1.74 +/- 0.02 1012#108.l 109 0.161 3.93 +/- 0.07 1.81 +/- 0.02 1011#143.l 143 0.121 4.12 +/- 0.07 1.90 +/- 0.02 1011#159.1 159 0.089 4.22 +/- 0.06 2.08 +/- 0.01 1012#195.l 195 0.091 4.35 +/- 0.06 2.20 +/- 0.02 1013#259.1 259 0.075 4.60 +/- 0.04 2.47 +/- 0.02 1008#295.l 295 0.090 4.76 +/- 0.05 2.52 +/- 0.02 1013#304.1 304 0.103 4.74 +/— 0.05 2.58 +/- 0.02 1023#403.2 403 0.100 4.86 +/- 0.05 2.76 +/- 0.02 1023#455.2 455 0.082 4.93 +/- 0.03 2.88 +/- 0.02 1023#508.2 508 0.096 4.99 +/- 0.05 2.97 +/- 0.02 1023#571.2 571 0.278 3.93 +/- 0.23 3.06 +/— 0.10 1023#623.2 623 0.205 3.73 +/- 0.15 3.23 +/- 0.07 the carbon-carbon properties (or thermal model). This is a desirable characteristic for additional materials (i.e. mica heater and ceramic insulation) in the experimental model. A similar insensitivity is demonstrated for the thermal properties of the insulation. 2-4.2 Effective Properties of Carbon-Carbon The properties estimated for the carbon-carbon composite are given in Table 2-4. The first column identifies the experimental case and the second column the initial 28 temperature. The third column is the estimated standard deviation of the analysis, which is defined as 1 I .I 2 0.5 O = [W2 2 (Yij_ 7.0)] (2-18) 1 = l j = l where N t is the total number measurements (from both sensor locations, N t = I + J) and N p is the number of estimated parameters. Variables I and J are the number of measure- ment times and number of sensors, respectively. The last two columns present the esti- mated properties with the associated confidence interval (calculated by PROPID and discussed below). Insight about the estimated properties is gained by plotting the properties as a func- tion of temperature; Figure 2-3 gives a plot of the one-dimensional properties of the car- bon-carbon composite as a function of the temperature, with the estimated confidence intervals. Note that the analysis assumes that the thermal properties were constant during an experiment, but varied between experiments; the initial temperature of the experiment is used to plot the properties. Using an F-test (Beck and Arnold, chapter 6, 1977) it is con- cluded that a second order (in temperature) model adequately represents the results. The equations determined for the properties with a least squares fit are k" 342 338 T40 186 T-TO 2 9 = . . — . 21 y, CC + (T3 ‘ T0) (T3 _ To) ( ) C" 141 265 T40 111T—TO 2 106 0 (P )CC — [. + . (T3—TO)_ . (T3—TOHX (22) where T0 = 31°C and T3 = 508°C are the minimum and maximum temperatures. By normalizing the temperature in equation (2-19) and (2-20) the coefficients in the describing equations have units equivalent to the respective thermal properties. A physical appreciation of the magnitudes is more easily realized in this form. The relationships are also shown in Figure 2-3. The thermal conductivities determined from the last two experiments are not used in the least squares analysis. The validity of the thermal properties estimated at temperatures of 571°C and 623°C is uncertain. The residuals and confidence intervals for these two cases are relatively large. When the set-up was disassembled after these experiments, the outer coating of the specimen (which protects from oxidation) had separated from the inner composite material. Because the estimated thermal conductivities are lower it is quite possible that the failure of the specimen 29 occurred just before or during these experiments. Thermal Conductivity [W/(mC)] 0 '100'200'300'400'500'600 Temperature (C) 10.01 . . . r . 3.5E+06 9.0- 5 - c " —3.0E+06 8.0- (p )“ . 7.0.. -2.5E+06 50': kg,“ -2.0E+06 5.0- . 450‘ :1.5E+06 30, -1.0E+06 2'0- .501: 05 1.0- _ ' + 0.0 0.0E+00 Figure 2-3 Temperature dependence of thermal properties estimated from one- dimensional experiments Volumetric Heat Capacity [J/(m3C)] 30 The details of the parameter estimation are quite similar at different initial tempera- tures. For this reason, and for brevity, only one case (1023#508.2), at an initial tempera- ture of 508°C , is discussed. The experimental data for this case are presented in Figure 2- 4. The temperatures shown represent the average of all thermocouples at the referenced locations. The results closely approximate the case of a finite slab heated with a constant heat flux at the surface and insulated at the backside. Initially, the temperature at the heated surface increases rapidly, while the backside temperature remains constant. Later both the temperatures at the heated surface and backside increase linearly with time until the power to the heater is turned off. Then the temperatures at both surfaces tend to approach the same temperature, demonstrating that there is little heat loss to the insulation (even though it is considered in the analysis). 525.0 520.0- 8 515.0- a a e g» 5100- .42 o E i” E 505.0- Heat Flux {15000.0 3 _-1oooo.0 a 35000.0 ‘5}; 500.0 - . . . 4 4 . , 0.0 m 0 10 20 30 time (seconds) Figure 24 Experimental data for one-dimensional case 1023#508.2 31 In addition to estimating the thermal properties, PROPlD provides some means to quantify the accuracy of the estimates. The previously discussed estimated standard devia- tion, equation (2-15), provides an indication of how well the calculated match the mea- sured temperatures. The magnitude of ("I can be compared to the temperature rise of the experiment, which is approximately 15°C . Except for the last two experiments, 6 is within 1% of the temperature rise. Also given with the estimates of the parameters is a confidence interval (Beck and Arnold, Chapter 6, 1977). The calculation of the confidence intervals has associated assumptions. First, the model for the experiment is correct. Sec- ond, the dominant errors in the analysis are in the temperature measurements, modeled with a first order auto-regressive model, and the errors are not biased (Beck and Arnold, 1977). Other quantities can be observed to demonstrate the accuracy of the estimated proper- ties. These include the sequential estimates of the properties, the residuals, and the sensi- tivity coefficients. The quantities are important to provide insight to the estimation, as well as insight to the experiment. Observing them can help improve the experiment and support the accuracy of the estimated properties. Each is discussed below. The sequential estimates demonstrate how the estimated properties vary as additional measurements are considered. The analysis assumes that the estimates are linearized about the converged parameter values using all experimental data. Figure 2-5 shows the sequen- tial estimates for this case. The sequentially estimated property, at time I, , represents the outcome if only data up to that time is used in the analysis (and linearized about the con- verged property values). In other words, if the data is analyzed by adding one data pair, Y( y = 0) and Y( y = Ly) , at each time, it shows how the estimated properties change as 32 6.0 . . . . . . . e 3.2E+06 :Q —\V’i (pC)CC : Lg A e -2.8E+O6 (,9 i 5.0" V— ky, cc g E L2.4E+00 a .§ 4.0- r? ,2 -2.0E+06 g g 8* g 3.0- -1.65+06 3 O a U 0 "a -1.2E+06 f E 2.0- '5 2 -a.01:+05 “é {-1 :r 1.0- .5 -4.0£+05 > 0.0 . r 0.OE+00 0 5 1'0 1'5 2'0 2'5 30 :55 40 time (seconds) Figure 2-5 Sequential parameter estimates for one-dimensional case 1023#5082 an additional data pair is added to the analysis. Initially the sequential estimates vary because there is not enough information to determine the parameters. However, as more data is considered, the property estimates approach constants and the linearization approx— imation is accurate. Meaning, if the experiment (or analysis) is ended at 15 seconds, the estimated properties would not differ significantly from the properties at 30 seconds. In general, for a good estimation, the sequential estimates converge to a constant and are fairly steady with time. For times greater than fifteen seconds, the sequential estimates of k; CC and (pC):c vary 0.9 and 3.5%, respectively. These values can be compared with the confidence regions predicted in Table 2-4 for this case of 1% and 0.67%, respectively. These sequential results for (pC):c at 508°C are not as accurate as the confidence inter- val in Table 2.4. Uncertainty in the other experimental measurements account for the dif- ference; the confidence intervals include error in the temperature measurements only. 33 The residuals are related to the estimated standard deviation and are calculated using and represent the difference between the measured and calculated temperature for a partic- ular time (ti) and sensor location (j) . The estimated standard deviation gives an indica- tion of the magnitude of the residuals; the signs and magnitudes of the residuals can provide considerable insight. Figure 2-6 presents the residuals for this case. The magni- tude of the residuals is approximately 0.1°C . The residuals are correlated; most of the residuals are positive during the heating. This outcome may signify that some inconsis- tency exists in the model or that a small effect was omitted. However, the magnitudes of these residuals are small, within 1% of the temperature rise during the experiment, indicat- ing that errors in the model are minimal. 0.5 l I I 1 l 1 I 0.3- — . .4 - '01 . -"' - A :I. ‘ l." V‘ \ ' . U 0°1- . I I "5 '1“: 1i - v II. It. .5 I: a .l I ~ I . WI. 1‘ q E t . 2'1: 3'” t r U) _ _ U s. c 3 . ' _ 32 0.1 I; 615...": .1 -0.3- —— sensor (y = 0) _ ----- sensor (y = Ly) ‘005 l l l r I I 051015 20 25 30 3540 time (seconds) Figure 2-6 Temperature residuals for one-dimensional case 1023#508.2 34 Sensitivity coefficients are the first derivative of the temperature with respect to the parameters, thermal conductivity and volumetric heat capacity. They are indicators of how well designed the experiment is. In general, the sensitivity coefficients are desired to be large and uncorrelated (linearly independent). A sense of the magnitude of the sensitivity coefficients is gained through normalizing the sensitivity coefficients by multiplying by parameters, resulting in units of temperature for the normalized sensitivity coefficients. A comparison is then possible with the temperature rise of the experiment. For a well designed experiment, with boundary conditions similar to the case investigated in this study, the sum of the normalized sensitivity coefficients for the thermal conductivity and volumetric heat capacity is nearly equal to the negative of the temperature rise (equal only if perfectly insulated at y = Ly ). Sensitivity coefficients are useful in the design of exper- iments, i.e., determining the heating and experiment duration, location of sensors, heating area (for two-dimensional case), etc. A study of the sensitivity coefficients, prior to per- forming experiments, can lead to better experiment designs. Figure 2-7 shows the sensitivity coefficients for the representative experimental case (1023#508.2). The sensitivity to the thermal conductivity and volumetric heat capacity are shown for both sensor locations. The magnitudes are about equal to the temperature rises, which is a good feature. Notice that the sensitivity coefficients are correlated (linearly dependent) for times up to 10 seconds for the sensor at the surface of the specimen (y = 0). This is similar to the situation that resulted in only being able to estimate one parameter for the effective properties of the mica heater in the analysis of the short dura- tion experiment. In this case, however, information is available from another sensor . . . . . e e - (y = Ly) where the normallzed sens1t1vrty coefficrents for k“ C and (pC)CC have qurte C 35 different shapes (i.e., not correlated). Even though the (pC);c normalized sensitivity coefficient at (y = 0) and L), are both negative and decrease with time, the k; CC normal- ized sensitivity coefficients are different shapes with one decreasing (y = 0) and the (y = Ly) values being positive. The difference in the k; cc and (pC):c normalized sensi- tivity coefficients at (y = 0) and Ly is more pronounced as time increases and also after 22 seconds when the power to the heater is turned off. These sensitivity coefficients for k; cc and (pC):c show that the experiment is well-designed because 1) the sensitivity . e . . e . coefficrents for ky, C are qurte different from those for (pC)cc and 2) the normalized C magnitudes are large (relative to the temperature rise of the experiment). I I I l J _ l 1 A 2" / I \ (y = Ly) _ e / ” a " I’ d c: . / . .2 U E " . —2~ - g» - (y = 0) - IE . . ...6— .- . e e (y = 0) . _ _ ky’ccaT/aky,“ . . —— (0C)ECBT/B(0C)§c (y = Ly) . _1G I I I I I 1 1 0 5 1 0 1 5 20 25 30 35 40 time (seconds) Figure 2-7 Sensitivity coefficients for one-dimensional case 1023#508.2 36 2-5.0 Experimental Uncertainty The estimated thermal properties are presented with confidence intervals indicating the error in the estimates due to errors in temperature measurements. Only errors in the temperature measurements are considered in such an analysis. Errors in other experimen- tal measurements are not included in the confidence interval. Several other experimental aspects influence the estimated thermal properties. These other experimental issues are not the same nature as the errors in temperature. They are systematic errors and not random in nature. The uncertainty in the estimate parameters (5), based on the uncertainty in the experimental parameters, is calculated as .. . A 1/2 ~ ab 2 ab 2 3b 2 db = —d + —d + . . . —d 322 {(5.21) 0.. .2) +0., ..1 < > where z = [z], z], . . . tsz are experimentally measured parameters. In the experiment to estimate the thermal properties of the carbon-carbon, uncertainties in the measured heat flux, location (depth) of the thermocouple, thickness of the carbon-carbon specimen, effective thermal properties of the mica heater, and effective thermal properties of the insulating material are considered. An experimental uncertainty in the thicknesses of the mica heat and insulation are not included. This is because effective thermal properties were experimentally estimated for these materials, implicit in the effective thermal proper- ties is the prescribed thickness of the material. Consequently, uncertainty in the material properties includes an uncertainty in the thickness. Experimental uncertainty for each experimental parameter and its contribution to the total uncertainty in the estimated parameters are given in Table 2-5.Uncertainty in the measured heat flux was computed as 37 Table 2-5 Experimental uncertainty for the two-dimensional experiments P3328“ Uncertainty Contribution (g—Edz) (pC):Cx10"° 14,“. z dz J/(m3°C) WI q 125 W/m2 0.033 0.049 L), 0.05 mm 0.016 0.027 y, 0.025 mm 0.014 0.015 kg,“ 20% 0.04 0.0 (pC)f,,,.ca 20% 0.06 0.06 kfm 20% 0.0 0.0 (pC)fm 20% 0.0 0.0 TOTAL 0.082 0.083 1 dq = {(gg-d/ijz + (3%;de +(%1d1j}2 (3-23) where uncertainties in measurements of area (A), voltage (V), and current (1p) are included. There is not a dominant term in the uncertainty analysis. All experimental conditions considered are of the same order. Overall, the uncertainty due to other experimental mea- surements are excellent. The uncertainty represents a maximum of 2.4% in k; CC and 5.8% in (pC):C for the experiment at the lowest temperature and 1.6% and 2.8%, respec- tively at the highest temperature. These values assume the uncertainties do not vary with temperature, which is reasonable for these experiments. Chapter 3 CARBON-CARBON THERMAL PROPERTIES: TWO- DIMENSIONAL EXPERIMENTS 3-1.0 Introduction In this chapter two—dimensional experiments are presented to estimate two compo- nents of the thermal conductivity and the volumetric heat capacity for the carbon-carbon composite. As in the previous chapter, effective thermal properties are estimated that rep- resent the carbon matrix carbon fiber composite with a silicon-carbide protective coating. The properties, k; CC, (pC):C, determined from the one-dimensional experiments are not imposed on the two-dimensional solution. All the thermal properties are estimated simul- taneously for the two-dimensional case. This permits a comparison between the results from one-dimensional and two-dimensional experiments to demonstrate the consistency of the methods. 3-l.1 Motivation Interest in the solution of multi-dimensional inverse problems has gained momentum, particularly in recent years. As the importance and wide-spread application of inverse methods are realized, so too have the demands on the complexity of the problems that can be solved. Such is the case for the application of inverse methods to the field of heat transfer. Two examples are the estimation of the thermal properties of a material and the 38 39 determination of the heat flux at a boundary, both from experimental measurements. The latter problem is the inverse heat conduction problem (IHCP), which has been the main focus of research on multi-dimensional inverse problems in heat transfer. The application of inverse methods to evaluate the IHCP or estimate thermal properties are closely related, however. A variety of methods which are used to solve the one-dimensional IHCP, have been extended to the multi-dimensional case. Osman and Beck (1990), Hsu et al. (1992) and Bass (1980) use methods based on the function specification method (Beck, et al., 1985). Murio (1993b) has presented a mollified space-marching algorithm. The adjoint method is employed by Jamy et al. (1991) and Truffart et al. (1993). Alifanov and Egorov (1985), Alifanov and Kerov (1981), and Alifanov ( 1994) have presented formulations for iterative regularization methods to solve the two-dimensional problem. The Monte-Carlo method was investigated by Haji-Sheikh and Buckingham (1993). Solving the multi-dimensional IHCP is further discussed in Chapter 4. Although much energy has been focussed on the multi-dimensional IHCP, less work has been afforded to the estimation of thermal properties using inverse methods for the multi—dimensional case. Loh (1991) performed an experimental investigation for the esti- mation of thermal properties, orthotropic thermal conductivity and isotropic volumetric heat capacity, in a carbon-carbon composite. J arny et al. (1991) formulated the analysis to estimate the thermal conductivity using a conjugate gradient method with an adjoint equa- tion. The lack of research on the multi-dimensional estimation of thermal properties may be due to the small number of materials that display an appreciable anisotropy. Due to the 40 construction and advancement of composite materials, however, anisotropic thermal properties are inherent in the composite; the magnitude of the anisotropy depends on the type of materials. For the carbon-carbon material investigated by Loh the thermal conductivity varied nearly an order of magnitude for directions normal and parallel to the fiber direction. This anisotropic nature of the composite material requires multi- dimensional inverse solution methods to accurately determine the thermal properties. Although this chapter focuses on a laboratory method, the extension of a method to the field, i.e. while the aircraft is on the runway or in the hanger, is of particular interest for the carbon-carbon material because of the variability that the thermal properties demon- strate. An in situ method also allows changes in the material properties to be tracked dur- ing development and operation of the vehicle. The methods presented herein are not easily extended experimentally to a field application, due to the practicality of instrumenting the material. What is demonstrated, however, is the applicability of the analysis and algo- rithms to determine the thermal properties given experimental data. More work on the design and optimization of the experiments is required to move the methods to the field. 3-2.0 Experimental Aspects The same experimental apparatus, shown in Figure 2-1, is used, except that only one of three heaters is energized. The two-dimensional thermal model is shown in Figure 3-1. The locations for the thermocouples are given in Table 2-1. The thermal model is mathe- matically represented as three coupled problems. Although the techniques used herein can be extended to temperature variable properties in a given experiment, the models given below are for thermal properties that are temperature independent. 41 11— o I o o O L n ceramic insulation " ms W 6" 1 1 1 1,13 . . 2.14 : ll / composrte specrmen ( L T/Ciifi 3,8 4,9 5,10 6,11 7,12 t y ’ 131 1:1 0 o o N A i I" / , 77/ 1" <<<<<<<<<<(((((K/<<<<<<<<((((. ' 11111100011011“ T x q = 0 (10) L . 1/2 mica ””0" heater L assembly Figure 3-1 Heat transfer model for two-dimensional experiments The first problem models the mica heater/contact resistance 2 2 k8 a Tmica a Tmica _ C e Tmica _Lmica <)’< O 0 mica —2 +—2 — (P )mica a ,t> 33’ 33‘ 0 x 8T —“ =0 OSySL 8x x=0L y TQ(L%0)=7},O0 (3-3) (3-4) (3-5) (3-6) (3-7a) (3-8b) (3-8c) (3-9) (3-10) (3-11) (3-12) 43 Taktak (1992) investigated optimal experimental conditions to estimate the orthotro- pic thermal conductivity for this geometry. However, his model specified a temperature boundary condition at the backside of the composite, instead of (an approximate) insula- tion condition which is used here. Taktak showed that for two sensors, the extremes (x = 0, 7.62 cm) on the heated surface (y = 0) , was optimal for the condition that heat- ing occurred over one-half of the surface. For this case, which is approximately insulated at the backside of the specimen, includes the thermal effects of the heater assembly, and heats over one-third of the surface, a similar outcome would be expected. Measurement of the temperature on the surface where the heater is active and locations where it is not active provides contrasting effects (sensitivity coefficients), which improves the accuracy of the estimates. See the optimality criteria in Section 2-3.0. 3-3.0 Analysis Procedures The same analysis procedures presented in Section 2-3.0 are used for the two-dimen- sional case. However, now a two-dimensional heat conduction thermal model is solved. Computer program PROP2D implements this inverse method to determine two com- ponents of thermal conductivity and the volumetric heat capacity. The program was devel- oped at Michigan State University by taking the finite element code TOPAZZD (Shapiro, 1986) and combining this direct problem solver with these parameter estimation methods, to create a powerful algorithm. PROP2D allows for the estimation of the thermal proper- ties for multiple materials, with possibly irregular geometries, from transient measure- ments. The thermal conductivity can be orthotropic and temperature dependent thermal properties are allowed. 3-4.0 Results and Discussion A separate, independent set of one-dimensional experiments (Chapter 2) were per- formed to determine the thermal properties of the mica heater and ceramic insulation in the model (Figure 3—1). Effective properties were determined to account for imperfect con- tact between the layers. Therefore, only the thermal properties of the carbon-carbon com- posite are unknown in the model (Figure 3-1). Furthermore, one-dimensional experiments e were performed to determined the thermal conductivity normal to the fibers (ky, cc ) and the volumetric heat capacity (pCzc) in the previous chapter. The one-dimensional results provide initial estimates for the two-dimensional analysis and permit for a comparison to demonstrate the accuracy and consistency of the methods; one-dimensional results are not used as prior information in the analysis. For the two-dimensional experiments, the analysis is more sensitive to the experimen- tal conditions. For example, the magnitude and duration of the heat flux must produce ade- quate response (sensitivity coefficients) for the sensors nearer the active heater, as well as for the sensors farther from the active heater. This requires a longer heating duration than that used for the one-dimensional experiments. The locations of the thermocouples must also be known accurately, especially on the active heater where temperature gradients are large. Although it is not too difficult to locate the position of the thermocouples in the car- bon-carbon specimen, it is difficult to align the mica heater assembly relative to the speci- mens since the heating elements are not visible in the heater assembly. The two-dimensional thermal properties of the carbon-carbon composite determined for temperatures up to 400°C are given next. Experiments were conducted at regular intervals over this temperature range and analyzed assuming the thermal properties were 45 constant for the duration of an experiment, but varied between experiments. The measured experimental data and details of the parameter estimation are presented and discussed for a typical experiment. 3-4.1 Experimental Data Experimental data for a test started at a temperature of 297°C (experiment case 1022$297) are shown in Figure 3-2. A sampling interval of 0.64 seconds is used to acquired data for this experiment. The heating begins at approximately 16 seconds and ends at 56 seconds. For experiments at lower initial temperatures the heating duration is 80 seconds. However, based on observation of the sensitivity coefficients and the criteria for optimal experiments, “D-optimality”, (Beck and Arnold, Chapter 8, 1977) a shorter 325°G""'I‘1v1u.. sensors 3,8 320.04 4,9 (5 315.0- E? § 510.0- 8. “E’ 3050 [— o 1 300.0- 6? :20000.0 § 295.0- Heat Flux ----- y = 0.9lcm _ ENE \ — y = 0.0cm FIOOOOD é a 290'0""'I'1rj-..'o,o§ 020"106080100120140 :15 time (see) Figure 3-2 Experimental data for two-dimensional case 1022$297, see Table 2—1 for sensor locations 46 duration, higher magnitude heat flux is selected to be closer to optimal experimental con- ditions for determining the two components of the thermal conductivity and the volumet- ric heat capacity. The complexity of this experiment is that two components of thermal conductivity and volumetric heat capacity are simultaneously estimated. The experimental conditions must be selected to provide information about all three effects. An alternative is to conduct an series of experiments, each experiment providing information on one (or more) particular effect. Then analyze the different experiments in a sequential manner. Osman and Beck (1991) used such a procedure to estimate temperature dependent thermal properties. For this model, a single experiment provides adequate information on all ther- mal properties (effects) and a sequential procedure is not required. The effect of the orthotropic thermal conductivity is seen by comparing in Figure 3-2 the temperature rise for the sensors at x = 3.81 cm (Figure 3-1) on the heated surface (sensors 5,10) and x = 1.27 cm on the insulated surface (sensors 1,13). The larger thermal conductivity in the x-direction results in a nearly instantaneous response at x = 3.81 cm on the heated surface, while the sensor at x = 1.27 cm on the insulated surface has approximately a four second time delay before responding. This delay exists even though the sensors are approximately the same distance from the active heater (~10% difference, 0.835 and 0.914 cm from the sensor on the heated surface and insulated surface, respec- tively). Notice that temperature data are acquired after the heating ends. Continuing to acquire data after stopping the heat flux results in better estimates. This is because it causes the sensitivity coefficients to be of different shape from one another after heating. These effects result in a more accurate estimation of multiple thermal properties based on 47 the criteria “D-optimality with constraints” (Beck and Arnold, pp. 459, 1977). Possible heat losses in the experimental set-up can also be monitored with this data, although it does not appear that there are significant losses in this experiment, since all temperature SCIISOI'S Converge IO a constant. 3-4.2 Estimated Thermal Properties Numerical issues are more important for the two-dimensional analysis than they are for the one-dimensional. The mesh size and time step selected for the finite element method can greatly influence the amount of time required to obtain a solution and the accuracy of this solution. The mesh used for this analysis contains 525 (quadrilateral) ele- ments. Twenty-five elements are along the 7.62 cm surface (x-direction) for all materials. There is one element across the mica heater assembly and ten elements across each the carbon-carbon specimen and ceramic insulation (y-direction). The computational time step chosen was 0.64 seconds, the same as the experimental time step. A typical two- dimensional analysis required 2-4 hours on a VAXstation 4000 Model 60 running VMS V5.5—2, depending on the number of iterations and accuracy of the initial parameter esti- mates; typically 4-5 iterations (16-20 direct solutions) were required. Although a detailed investigation was not performed, the time step and mesh size were varied and shown to have little influence on the resulting estimated thermal properties. Two-dimensional thermal properties determined for the carbon—carbon composite are summarized in Table 3-1. The experiment case number and initial temperature are given in columns one and two. The next four columns present the estimated standard deviation (6) and the two-dimensional thermal properties determined from the analysis with confidence intervals. The final two columns give the duration and magnitude of the applied heat flux. 48 Table 3-1 Thermal properties estimated for the carbon-carbon composite from two- dimensional experiments and analysis EXP T1812: A k; cc k; cc (P C):cx 10-6 Case p 6 Heat Flux (°C) (°C) W/mC W/mC J/m3c sec W/m2 1020&65 65 0.168 58.4 +/- 0.34 3.89 +/. 0.04 1.52 +7- 0.003 80 5276 10209111 111 0.183 61.0 +7- 0.39 4.17 +/. 0.04 1.74 +/- 0.004 80 6095 102119.157 158 0.163 60.7 +/- 0.34 4.24 +/. 0.04 1.93 +/. 0.003 80 6910 1021&205 205 0.210 61.8 +/. 0.33 4.55 +7- 0.04 2.13 +/. 0.004 80 8905 1021&256 256 0.202 61.6 +/. 0.30 4.73 +/. 0.05 2.34 +7. 0.004 80 8792 10228297 297 0.295 58.8 +/. 0.38 4.97 +/. 0.06 2.36 +/. 0.007 40 17304 10223352 353 0.277 58.7 +/. 0.37 5.07 +/. 0.06 2.56 +/— 0.007 40 17259 1022$403 403 0.256 57.4 +/- 0.34 5.09 +/- 0.05 2.66 +/- 0.007 40 17096 A plot of the thermal properties (ke kg and ( p C )2.) with estimated confidence x, cc ’ y, cc ’ interval as a function of the initial temperature is shown in Figure 3-3. A F—test (Beck and Arnold, Chapter 7, 1977) indicated a second order model (in temperature) for these prop- erties. Because limited results are available, the relationship for k: cc remains linear. The relationships determined with a least squares fit, are shown in Figure 3-3 and given by the following relationships: kg 387 193 LT‘ 0656 T_T' 2 = . . —. -13 e T— l kt,“ = 60.8—2.07(T T J (314) 2_ 1 Ce 152 168 LT‘ 0543 FT‘ 2 10° 3 5 9-1-1 1—1 11x C TZ—Tl TZ-‘Tl where T1 = 65°C and T2 = 403°C. The thermal conductivity parallel to the fibers 49 8.0 . . . . - . - . . 8.0E+06 7.0 - e - 7.0E+06 _ . ky, cc .. 6 66.0- 2T1: =2: _._ ~6.0E+06§ é - = . E E 5.0 — - 5.0E+06 .§‘ 0 3? “ e ‘ 8. E 4.0 _ kx, cc '1 4.0E+06 6 8 - « a "O Q) g 3.0 - - 3.0E+06 I o . . .9. "a 5 2.0 — - 2.0E+06 E . 9C; , g ..C: O 1‘ 1.0 - - 1.0E+06 > 000.0 5 100 ‘ 200 ‘ 300 ‘ 400 L 50(9-0E+O6 Temperature (C) Figure 3-3 Temperature dependence of estimated thermal properties from two- dimensional experiments e y, cc (k. x, cc) is 12 to 15 times as large as that normal to the fibers (k ). Testing was halted at 400°C due to the failure of the carbon-carbon specimen during subsequent one-dimen- sional testing at higher temperatures. In addition to estimating the thermal properties, PROP2D provides some means to quantify the accuracy of the estimates. The estimated standard deviation, 6' computed in equation (2-18), which is given in Table 3-1, provides an indication of how well the calcu- lated temperatures match the experimentally measured temperatures. The magnitude of the estimated standard deviation can be compared to the temperature rise of the experi- ment, which is approximately 20 to 25°C. The magnitude of the estimated standard devi- ation is within 1.2% of the maximum temperature rise for all the experiments and many 50 are less than 1%. The estimated standard deviation is largest for the last three experiments, which applied a larger heat flux for a shortened duration. There are other quantities that can be observed to demonstrate the accuracy of the estimated properties. These quantities are the sequential estimates of the thermal proper- ties, the temperature residuals, and sensitivity coefficients. These quantities are sensitive indicators to provide insight to the estimation and experiment. Each is discussed below for experiment 1022$297. The sequential estimates demonstrate how the estimated properties vary as additional experimental measurements are considered. Figure 3-4 shows the sequential estimates for this experiment. A sequentially estimated property, at time ti, represents the outcome if 70.0 . l . . . . . . . . . 3.0E+06 A 50.0“ __ I‘M; ~2.5E+06 5; (90).. \ 50.0- E -2.0£+06 :5 40.0- % -1.5E+06 g 30.0- D a -1.0E+06 g 20.0- 5 1 00- e '5.0E+05 —V¥ ky, CC 000 I T I I U 0.0E‘I'OO time (sec) I'jTl'l'l‘l 0 20 40 60 80100120140160 Figure 3-4 Sequential parameter estimates for two-dimensional case 1022$297 Volumetric Heat Capacity (J/m3C) 51 only data up to that time is used in the analysis and the results are linearized about the con- verged parameter values. In other words, if the data is analyzed by adding one data set at each time, it shows how the estimated properties change as one more data set is added to the analysis. Initially the sequential estimates vary because there is not enough informa- tion (data) to accurately determine the properties. However, as more data is considered, the property estimates approach constants. If the experiment (or analysis) is ended at 80 seconds, the estimated properties would not differ significantly from the properties at 100 seconds. In general, for a good estimation the sequential estimates converge to a constant and are steady with time. For this experiment the sequential estimates for k; cc, k; cc , and (pC):C vary only 7.7, 3.0, and 1.6% over the final one-half of the experiment. The values are quite large compared to the confidence intervals, which were 0.6%, 1.2% and 0.3%, respectively. The confidence interval models error in temperature only. Uncertainty in other experimental measurements account for the discrepancy. The temperature residuals are related to 6 and are calculated as shown in equation (2- 18). They represent the difference between the measured and calculated temperature for a particular time (ti) and sensor location (x j, yj) . The estimated standard deviation gives an indication of the magnitude of the residuals; the sign and magnitude of the individual residual can provide considerable insight. The residuals for this experiment (1022$297) are shown in Figure 3-5a, b and c. Figure 3—5a presents the residuals for the sensors on the active heater, Figure 3-Sb the residuals for the sensors on the heated surface but not on the active heater, and Figure 3-50 the residuals for the sensors at the insulated surface. The residuals for the sensors on the heated surface, but not on the active heater (Figure 3-5b), are slightly correlated. The other residuals, Figure 3-5a and 3-50, are highly correlated and 52 Residual (C) o $ —0.504 ----- Sensors 4,9 _ —1.00- Sensors3,8 -1.50r1........ . a 1 ' 0 20 40 60 80 1 00 1 20 1 40 time (sec) Figure 3-5a Temperature residuals for two-dimensional case 1022$297 on surface below active heater 53 1.50 I | I I ' I ' l 1 T ' I ' I 1.00- _ ,.. 0.50- - 8 . ' . . -— ' 1 " I1 . , i ' \ . \ [Is I“: n d .’ g 0.00.. I 0* ‘ W '1 1 l r b 4 «1‘3 . 1 . m I .0501 _ Sensors 7,12 _ " _____ Sensors 6,11 ‘ _1‘OO- -— Sensors 5,10 3 . 4 “'1 .50 I I I I I I l I r I I I 1 O N O ‘5 O) O a: O —I O O a... N O E .4. time (sec) Figure 3-5b Temperature residuals for two-dimensional case 1022$297 on the heated surface, but not on the are of the active heater 54 1.00- 0.50- g ., 'r T; 0.00- ."2 ~ . , E I “‘0' Io‘,"1't.’.".‘t"0'-“o‘ ‘1‘ 0" 51v- ‘3’! -0.50J . ----- Sensors 2,14 -1.005 '4 Sensors 1,13 '1°50""'Iru-i..., 0 20 40 60 80 100 120 140 time (sec) Figure 3-50 Temperature residuals for two-dimensional case 1022$297 at the specimen/insulation interface 55 larger, 4—5% of the temperature rise on the heated surface and 8-10% of the temperature rise at the insulated surface. It is not clear exactly why the residuals are so highly corre- lated. Two possible reasons are the uncertainty in the location of the sensors with respect to the heater and a non-uniform heat flux produced by the heater. First, there is some uncertainty in the alignment of the heater assembly and the specimen. The heating ele- ments are contained within an opaque fiberglass with a mica layer on the outside. Since the fiberglass and mica extend beyond the heating elements, the alignment of the edge of the heater (element) with the specimen is difficult. Second, the design of the heating ele- ment has a gap between successive coils allowing for areas of localized heating. If the sen- sors align with a gap, the actual heat flux will be larger than that calculated flux, which assumes a uniform profile. Similarly, the sensors aligned with the heating element will have a larger heat flux than the calculated uniform flux. The residuals for sensors away from the active heater are less sensitive to the location and uniformity of the heat flux and therefore are less affected. To investigate the correlated residuals, two numerical experiments were conducted. The first numerical experiment investigated an error in the location of the heater and the second investigated a non-uniform heat flux; no other measurement errors were present. These experiments were analyzed assuming no error in the location of the heater and a uniform heat flux to investigate the effect on the residuals. Unfortunately, the results of these numerical experiments were not conclusive. In particular, the two errors had opposite effects on the residuals. An error in the location of the heater resulted in a residual that had an opposite sign to the residuals for a non-uniform heat flux. The shapes of the residual curves for the numerical experiments were quite different from the 56 2 experimental case (Figure 3-5), especially for sensors not at the location of the applied heat flux. As previously noted, observation of the sensitivity coefficients can provide insight to the estimation problem. An advantage of parameter estimation, compared to other inverse methods, such as gradient methods (J amy et al., 1991), is that the sensitivity coefficients are computed in the analysis. Hence, no additional computation are required to observe the coefficients. Observation of the sensitivity coefficients at this stage may be too late, since the experiment is essentially designed and moving sensors or changing the heated area is not easily done. However, some minor modifications may improve the accuracy, such as changing the heating duration or magnitude. When possible, an analysis of the sensitivity coefficients should be conducted prior to running the experiments. Sensitivity coefficients provide insight to design experiments. In general, the normal- ized sensitivity coefficients are desired to be large for each parameter and uncorrelated (linearly independent) for different parameters. A sense of the magnitude of the sensitivity coefficients is gained through normalizing the sensitivity coefficients. Normalization is performed by multiplying by the parameters, resulting in unitsof temperature for the nor- malized sensitivity coefficients. The normalized sensitivity coefficient for parameter 1] is — 8T X11 = n- (3-16) A comparison is then permitted with the temperature rise of the experiment. For the cur- rent experimental design, the normalized sensitivity coefficients are shown in Figure 3-6a- c. Figure 3-6a shows the sensitivity to pCZC , ch, and Figure 3-6b and c shows the sensi- tivity to the thermal conductivities k: c .C and kin, X,“ and ka. Because the sensitivity 57 coefficients are normalized, a direct comparison of their magnitudes is possible. Some observations are drawn from the sensitivity coefficient plots. The most information is available on the active heater (sensor locations x = 0.89 cm and 1.91 cm for y = 0). At these locations the sensitivity coefficients have the largest magnitudes and therefore are most influential on the values of the estimated properties. Notice that Xky undergoes a sign change across the specimen (in the y-direction) and ka 6 I r I I I I 4_ - g 2- ,,,,,,, y=0.91cm ‘ B C. d) '8 O - 35 O 0 U _2- ‘ 3‘ IE .2 _4_ - d.) m -o -6, 1 ”<3 5 _3r _ t Z -‘ ‘10— sensor 3,8 I 1,13 2,14 d —12 ' ' ‘ ‘ ‘ L 0 20 40 60 80 100 120 140 time (seconds) Figure 3-6a Normalized sensitivity coefficient for volumetric heat capacity, X pC, for two-dimensional case 1022$297 58 6 I . 1 I I I - 1,13 _ 4 . ......... y=0 A ....... =0.91cm B 2- y _ a 712 2,14 0 0 ............... . ...... q E (3 6,11 3 '2" - 35 5,11 2 _4_ I <1) m B .5 '6“ _ E“ o -8— q Z sensor3,8 —10- - -12 L l 1 1 1 1 0 20 40 60 80 100 120 140 time (seconds) Figure 3-6b Normalized sensitivity coefficient for thermal conductivity in y-direction, 2,), for two-dimensional case 10225297 Normalized Sensitivity Coefficients (C) -10 —12 0 Figure 3-6c Normalized sensitivity coefficient for thermal conductivity in x-direction 59 sensor 3,8 y 0 ....... y - 0.9lcm . 20 40 60 80 100 120 time (seconds) X k! for two-dimensional case 1022$297 140 60 undergoes a sign change along the specimen (in the x-direction). The implications of the these sign changes are that there exist 1) a y-location within the body where the tempera- ture is insensitivity to ky’ CC and more importantly 2) a x-location where the temperature is insensitivity to k x, CC. The latter result is more important for this case where surfaces tem- peratures are measured, because seemingly logical locations along the measurement sur- faces y = 0, Ly may be insensitive to kx, CC. Although it is desirable to avoid locations where the sensitivity coefficients changes sign because the sensitivity is quite small, sen- sor locations that have sensitivity coefficients with opposite signs are a beneficial result. This situation produces contrasting effects which improves the accuracy of the estimates. Hence, the choice of the locations near the edges of the specimen (x = 0, 7.62 cm), which have oppositely signed sensitivity coefficients, for the measurement locations. For— tunately, the surfaces of the specimen (y = 0, Ly) are the most sensitive in that direction. Finally, the experiment was halted after heating for approximately 40 seconds to make comparable the magnitudes of the sensitivity coefficients for all thermal properties. In Fig- ure 3-6b 70.) is approaching a steady-state value, while the sensitivity to X k! and XPC have not. The sensitivity coefficient ch does not have a steady-state solution and will continue to increase linearly with time. If the experiment heated for a longer duration Xk)‘ and ch would be much larger than X 1;,- Consequently, the estimates for the two former properties would be considerably more accurate than the latter property. 3-4.3 Comparison of One- and Two-Dimensional Results The two-dimensional results can be compared to the one-dimensional results in chap- ter two for the thermal conductivity normal to the fiber and the volumetric heat capacity. The results obtained for the one-dimensional analysis were used as initial estimates for the 61 two-dimensional analysis; however, the thermal properties were not constrained in the two-dimensional analysis using the one-dimensional results. The properties determined from the one- and two-dimensional experiments are com- pared in Figure 3-7a and 3-7b. Overall, the results are quite close. The two-dimensional and slightly lower for (pC):c. Recall that the pr0per- C results are slightly higher for k; C ties were estimated assuming they were constant over the duration of an experiment and therefore are applicable for a 20-25°C temperature range. Noting this, the one and two- dimensional estimates for k; CC are within 6%, with the largest errors at the higher tem- peratures where limited two-dimensional data is available. There is also a dip downward for the estimates near the maximum temperature, an unexpected result that may indicate the thermal conductivity is approaching a constant, but more testing (at higher tempera- tures) is needed for verification. The results for (13C); have a maximum of 7% deviation between the one and two-dimensional estimates. The temperature dependence determined using both one- and two-dimensional results, which were also indicated to be second order in temperature using a F-test, are kg = 346+359 —208 3 ‘7 I I O -1 y'c (13 10) (7 3 10) ( ) Ce =ll4l+242 —08 '10 3 8 O O C -1 (P )CC (T3 7) 94( 3 ) x ( ) The relationship for k: CC has only two-dimensional results and is given by equation (3- 12). The estimated thermal properties, from both one- and two-dimensional experiments, are within 4% and 6% of the relationships given in equation (3-17) and (3-18) for k: CC and (pC):C. 62 2 O— . 2DExp O lDExp Thermal Conductivity (W/mC) ' ' 1 1 1 1 1 ' 1 O 1 00 200 300 400 500 Temperature (C) e 11) (PC )Cc 3.0E+06— 2.5E+06- 2.0E+O6-— 1.5E+06~ 1.0E+06— 0 21) EXP - « O 11) Exp - 5.0E+05— - Volumetric Heat Capacity (W/m3C) 0.0E+00 v 1 1 1 0 100 200 I I I I 1 1 1 1 300 400 500 600 Temperature (C) Figure 3-7 Comparison of thermal properties estimated from one- and two- dimensional experiments a) thermal conductivity k: CC and b) volumetric specific heat (0C); 63 Table 3-2 Experimental uncertainty for the two-dimensional experiments Paigeter Uncertainty Contribution (gig-dz) (pC1:.x10“° k2... k5,... z dz J/(m3C) W/(mC) W/(mC) q 3%*q W/m2 0.086 1.9 0.15 Ly 0.05 mm 0.014 0.43 0.066 y1 0.025 mm 5.5 13-04 0.14 0.016 x J. (i=1,2. J) 1.0 mm 8.0 13—04 0.72 0.20 kg,“ 20% 0.018 0.55 0.073 (003,“, 20% 0.020 1.2 0.14 kfm 20% 0.0 0.0 0.0 (0C); 20% 0.0 0.0 0.0 TOTAL 0.091 2.5 0.30 3-5.0 Experimental Uncertainty Using the techniques described in Section 4-5.0 for the one-dimensional properties, the experimental uncertainty for the two-dimensional thermal properties is assessed. For the two-dimensional experiment uncertainties in the measured surface heat flux, thickness of the carbon-carbon specimen, y-location and x-locations of sensors on the surface near- est the heater (experiment is insensitive to locations specified for sensors on back surface of specimen), thermal properties of mica heater, and thermal properties of insulation are considered. As in the one-dimensional analysis, the uncertainty in the thicknesses of mica heater and ceramic insulation are not considered because they are implicit in the effective properties estimated for these materials. Table 3-2 gives the uncertainties in the estimated 64 thermal properties due to the uncertainties in the measured quantities. For the two-dimen- sional case there are three important uncertainties. The uncertainty in the measured heat flux, x—location of the sensors, and the thermal properties of the mica heater are dominant terms. All terms contribute significantly to the overall uncertainty, but vary depending on the thermal property. The heat flux is important because there is a large uncertainty in the area associated with locating the heater relative to the sensors. The x-location of the sen- sors are important due to alignment issues. The two-dimensional uncertainties are larger than one-dimensional results. However, uncertainties are still a maximum of 6.0, 4.2, and 7.7% of(pC)" k" and kf’ CC , x, CC, v, CC, respectively. Chapter 4 SOLUTION OF IHCP USING A GRADIENT METHOD WITH ADJ OIN T EQUATION APPROACH 4-1.0 Introduction Estimating the conditions at the surface of a conducting body from internal measure- ments is typically called the inverse heat conduction problem (IHCP). Inverse describes this type of conduction problem because conditions at the boundary or surface of a body are estimated using internal measurements. Whereas a direct conduction problem, some- times called just the direct problem, uses conditions specified on the boundary to compute the internal temperature. While the direct problem is generally a well-posed problem, the inverse problem tends to be ill—posed and very sensitive to measurement errors. A sequen- tial method to solve the IHCP is discussed in this chapter. Applications of the IHCP are abundant. A classic problem is determining the heat flux at the surface of space vehicles during re-entry. A related problem is the estimation of the heat flux at the leading edges of hypersonic vehicles. Other examples are estimating the heat flux, or convection coefficients during quenching, which enables control in the result- ing material properties. During casting processes, estimates of the heat flux can be used to the control the cooling rate and movement of the cooling front to minimize residual stresses and void formation. Measurements on the outside of the cylinder wall of a internal 65 66 combustion engine can estimate the heat flux in the combustion chamber. Additional examples are cited in Kurpisz and Nowak (1995). Several methods are applied to solve the IHCP. Function specification (Beck et al., 1985), Tikhonov regularization (Tikhonov and Arsenin, 1977), gradient methods (Ali- fanov, 1994 and Ozisik, 1993), and mollification (Murio, 1993) are more frequently cited methods. However, other approaches also applied are, dynamic programming (Busby and Truijillo, 1985), Kalman filter (Tuan et al., 1996), Monte Carlo method (Haji-Sheikh and Buckingham, 1993). In addition, combining methods is useful. Beck and Murio (1986) formulated a combined function specification and Tikhonov regularization method, simi- lar concepts were used in Osman et al. (1997). Jamy et al. (1991) combined an iterative search method with Tikhonov regularization. Since this dissertation focuses on multi- dimensional problems, the literature review will narrow its attention to the two-dimen- sional problem. Refer to books by Alifanov (1994), Beck et a1. (1985), Hensel (1991), and Kurpisz and Nowak (1995), for a comprehensive survey of the literature. Many methods applied to the one-dimensional problem have been extended for the two-dimensional problem. Function specification and gradient methods have received the most attention. Function specification specifies a functional form for the heat flux over a future interval (Beck et al., 1985). Specifying a functional form over the future interval provides regularization to stabilize the ill-conditioned problem (Lamm, 1995). In conjunc- tion with specifying a function form, function specification solves the problem in a sequential manner. Several researchers have investigated the two-dimensional application of the function specification method. It was applied to estimate spatially and time varying convective heat 67 transfer coefficients, Osman and Beck (1989, 1990), and surface heat flux, Osman et a1. (1997). A boundary element method was coupled with function specification to investigate multi-dimensional problems by Zabaras and Liu (1988). Hsu et al. (1992) applied a finite element method to solve the general two-dimensional problem with inverse methods simi- lar to function specification. Gradient methods, which typically apply a conjugate gradient iterative scheme, cou- ple iterative or Tikhonov regularization to stabilize the solution and solve the multi-dimen- sional problem. Iterative regularization depends on the slowness or “viscosity” of the solution and uses the iteration index as the regularization parameter. Several papers use iterative regularization; see Alifanov and Kerov (1981), Kerov (1983), and Alifanov and Egorov (1985). Additional investigations using gradient methods, but not iterative regular- ization, are given in Zabaras and Yang (1996), Reinhardt (1996a.b) and Jamy et al. (1991). Others have studied the multi-dimensional inverse problem with a variety of approaches. Tuan et al. (1996) uses Kalman filter to develop an on-line algorithm. A new method, called “direct sensitivity coefficient,” is claimed by Tseng and Zhao (1996) and Tseng et a1. (1996). Pasquetti and Le Niliot (1991) employ Tikhonov regularization with the boundary element method. An adjoint equation approach to compute sensitivities and relate measured temperature to unknown surface conditions is used by Hensel and Hills (1989). A Monte-Carlo method is given by Haji-Sheikh and Buckingham (1993). Mollifi- cation with a space marching technique is used by Murio (1993) and Guo and Murio (1991). Busby and Trujillo (1985) use dynamic programming. An analytical solution is developed by Mosaad ( 1995) and transform methods are used by Imber (1974, 1975). 68 Function specification and gradient methods are popular approaches. An advantage of the function specification method is that the problem retains the causal nature, represented with a Volterra operator (Lamm, 1995), and can be solved sequentially with possible sav- ings in computational time and memory. The method, however, requires assuming a priori information about the (unknown) function. Gradient methods require no a priori informa- tion but typically solve for a function on the whole time domain, not taking advantage of the causal nature of the problem. The demonstrated success of gradient methods and the efficiency of a sequential implementation suggest a sequential implementation of a gradient method would be a powerful combination. A method that sequentially implements a gradient scheme, using an adjoint equation approach, is proposed and to be developed; additional stability is intro- duced by including Tikhonov regularization. The method is anticipated to benefit from not requiring a prescribed functional form, which is particularly useful on space, and the com- putational aspects of a sequential implementation. Other researchers who have proposed a sequential implementation are Reinhardt and Hao (1996a,b) and Artyukhin and Gedzhadze (1994); however, past implementations have addressed the one-dimensional problem with a very limited investigation of the method. This dissertation addresses the one- and two-dimensional problems for the sequential implementation. Although applying the sequential gradient method to linear one-dimensional problems is probably not needed or recommended, addressing the one-dimensional problem provides valuable insight to the method. This is the first known two-dimensional implementation of a sequential gradi- ent method and the most comprehensive, to the best of the author’s knowledge, study of the gradient methods. 69 Derivation of the equations to solve the IHCP using a conjugate gradient search method with an adjoint equation approach are presented in this chapter. Gradient methods that apply an adjoint equation approach have been widely discussed in the literature. In particular, the Russian community has championed these methods (Alifanov, 1994). How- ever, the mathematical basis of the method seems to have limited application of the meth- ods (at least for engineers), more so in the USA than elsewhere. Reasons for the limited use of the method in the USA can be partially attributed to the strong advocacy of the FSM there (Beck et al., 1985). However, more recently the power of the methods, particularly for multi-dimensional problems and problems with many estimated components in the author’s opinion, has been realized. In particular, Ozisik (1993), is an advocate of gradient methods. In addition, the methods have been presented in a less mathematically rigorous form (Jamy et al., 1991, Lamm, 1990, and Jamy and Beck, 1995). Section 4—2.0 to 4-6.0 develop the describing equations to solve the general multi- dimensional IHCP. Equations are developed in differential form and presented in the form of an algorithm in Section 4-7.1 for a whole domain solution. A sequential implementa- tion is discussed and an algorithm is presented in Section 4-7.2. Derivations up to this point in the chapter assume that the unknown function resides in a prescribed function space E, which is an infinite-dimensional space. Restricting the solution to a finite-dimen- sional space, Rn , is addressed in Section 4-8.0. The chapter is concluded with a section to summarize the chapter and provide some insight to the method. 70 4-2.0 Problem Statement of the Multi-Dimensional IHCP A schematic of the general multi-dimensional IHCP is shown in Figure 4-1. In the analysis that follows, the conduction problem is assumed linear, i.e., thermal properties do not depend on temperature. Extending the gradient methods for non-linear problems is discussed in Artyukhin (1996) and Loulou et al. (1996). For the sequential implementation to be developed, it is possible to temporarily linearize the problem and not consider it non- linear. Assuming the thermal properties are independent of temperature, the problem is mathematically formulated as follows: V. (k(r)VT(r, t)) = pC(r)[—aa—IT(r, t) + v - VT(r, 0], at)": :nSgt:) (4-2.1) _k, 97170", t) + h, T(r, t) = fi(r, t), (r) on 1“,, (i =1’ 2’ 3) (42.23) (t0—2Y,-0>1 u [T(dj, m + 114(1) — T(d,» (1)1}dt j=l 2 2 9 — ' ’ A 1 A ’ “Lialeflr { [610 I) qp,,(r 0111 610 0+1” q(r 1)] }dr W435) u The resulting change in the temperature caused by a variation of magnitude ItAq in the heat flux is related to the directional derivative of the temperature. The directional deriva- tive of T at q in the direction Aq, evaluated at (r, t) is DAan, t;q) = lim T(r’ “‘1 + “A” ‘ T(r, ”1) (43.6) u-—)0 [1 Taking the limit of equation (4-3.5) as It —> 0 while using the definition of the directional derivative of temperature in equation (4-3.6) gives 75 J DAquq> = 2 J2me. t;q) - Y ,(r)1DA,T(d,-. t;q)dr j: 1 + aTJ’fI [q(r, t) — qpnrr, t)]Aq(r, t)dr (1:14-37) to 1‘4 where the left side of equation follows from the definition in equation (4-3.2). This equa- tion shows that the directional derivative of the J (q) is a function of the directional deriv- ative of the temperature D A qT(d j, t;q) . The directional derivative of temperature, i.e., the variation in temperature due to a variation in the heat flux, is typically referred to as the sensitivity function and defined as 0(r, t;Aq) E DAqT(r, t;q) (4-3.8) The sensitivity function 0(r, t;Aq) represents the variation in temperature T(r, t;q) due to the presence of an input heat flux function Aq. 4-4.0 Sensitivity Equations The describing problem for the sensitivity function is derived from the definition of the directional derivative in equation (4-3.6). Evaluating the direct problem in equation (4- 2.1 and 4-2.2) at q + ItAq gives V ' (k(r)VT(r, I:q+11461)) = pC(r)[§—T(r, t;q + ItAq) + v - VT(r, t;q + ItAq)], (r) in Q (4-3.9) at (to < t S If) (r) on 1",, (i =1, 2, 3) (t0 0 of equation (4-3.l 1) and (4-3.12) while using the definition in equation (4-3.6) and the definition of the sensitivity function 0(r, t;Aq) E DAqT(r, t;q) , the describing problem for sensitivity function is V-(k(r)V0(r,t)) = pC([)[—a—0(r,t)+v - V0(r,t)], (r) i“ 9 (43.13) at (t0+ f jv(r.t){V - (k(r)VT)—pC(r)[3-t +v - VT]}dr 4: [On 3 + 2 Eryn”, ”(19. %%_hiT+fi(r, t)}lr dt ”r I31 ,‘ + I? In4(r, t)(k(r)g% + q(r, 1))dr dt + I710(r)[T(r, 0) _ Tom] d, (4.43) 01-44 Q Applying Green’s theorem and integrating by parts on time, the first integral in equation (4-4.3) can be rewritten as 79 JII‘IWO', t){V . (k(r)VT) - pC(r)[%lt~ + v . VT]}dr dt = 0:: £’I{V . (k(r)V\V) + [pC(r)g—}V+ v - V[pC(T)Wl]}T 61" d’ 052 + ffl‘l’kUg—g— “HT—313w dt-LIIPCUNITU . fildr dt 0F 0F ‘jPCWT )1?= tow-(4-4.4) n where F = F1 u F2 u F3 U F4. Substituting this integral into the Lagrangian in equation (4-4.3) gives 4111.11.40.14) = J(T.q>+f’j{V~1kdr dt— #ch >1;= ,Odr + 2]: 111,0: )k,[g% h T+f.,(r t)]dr dt+J:: [1140, t)[k(r)gI-fi +q(r 0].), 4, i=1 + I;o(r)[T(r, 0) - T0(r)] dr (4—45) 9 The Lagrangian has now been rewritten such that the differential operators in the first inte- gral apply to the Lagrange multiplier 01, instead of temperature T. The directional deriva- tive of the Lagrangian function A, for fixed multipliers, is SAW, 11;" A.» T, q) = DATAW, 111, AG, T, q) + DAqu, ‘11, 10, T, q) (4-4.6) where D ATA( w, 11,-, 10, T, q) is the directional derivative at T in the direction AT evaluated at (11!, n i, 21.0, q) and D A qA(w, m, 3.0, T, q) is the directional derivative at q in the direction Aq evaluated at (111, T1,, 71.0, T). An analogous expression, similar to equation (4-4.6), can be written for the directional derivative of J(T,q). These directional 80 derivatives can be obtained by using a similar methodology to that given in equation (4- 3.2). The directional derivative of the Lagrangian criterion for fixed multipliers is SAW, 11,-, to, T. q) = 51(T. q) + Jj’j{V - (k(r)V\v) + pC(r)? + v - leC(r)w1}AT dr dt on B 8 . t + [Thane—fin — k(r)AT3—¥]dr dt — £’IpC(r)wAT(v - n)dr dt — ({pCWAT )1]= I0dr + 2f: [11,0 t)l:k-aan —AT+— —h.AT]dr 811+]: [1140, t)[k(r)aan —AT,+Aq(r t)]dr d, i=1 + jl,(r)1AT(r, 0)] dr (44.7) 9 The solution of interest is for the case that the temperature satisfies the direct problem, i.e., T = T(q) . Hence, for this case AT 2 0 in equation (4-4.7), therefore the boundary con- ditions on sensitivity problem, equation (4-3.l 1) and (4-3.12), can be used to simplify this equation. Introducing the boundary and initial conditions from the sensitivity problem, the Lagrangian criterion in equation (4-4.7) simplifies to 6401!. T14). q) = 611T.4>+f’1{V- (k(r)Vw)+ 96013;" + v - V10Cw1}e dr 4: ()Q 3 2130— —JJ I:I{1I/Aq(r, ’)+[k(5)a-7 +pC(r)(v n)\|l]0 }dr dt i=1 -ij(r)(\1I8 )1, dr (4-4.8) Q f where, (i = 1) 13C! = K’Hkm 32—pC(r)(v-n)0]wdr dt (449) (Ir 1 81 (i¢ 1) IBC, = JTJI— k(r)%%’+[hi—pC(r)(v~fz)]w]6dr dt (4-4.10) ()ri Notice that the criterion is no longer a function of the multipliers “i and lo since the terms for these multipliers are zero from the boundary conditions on the sensitivity prob- lem. It can be shown that 51(T(q), q) = DAqJ(q) (4-4.ll) Defining the residual for the difference between the measured and calculated temperatures (ej(t) = YjU) - T(d,» I, 4)) in the directional derivative DAqJ(q) , equation (4-3.7), the directional derivative of the Lagrangian is SAW, T(q), q) = J _ 2 £II[ej(t)5(r—dj)]6 dv dz + an’jr [q(r, z) —qp,,.(r, t)]Aq(r, z) dr dz j=l 09 t0 4 +ij [{V - (k(r)Vw) + pC(r)-git" + v ~ V[pC(r)w]}9 dr dt "Q 3 + Z IBC, —J:f I{qu(r, t) + [k(r)g—:i: + pC(r)(v - ii)\|l]6} dr (1! or4 i=1 -IPC(f)(W9 )|t=t dr(4-4.12) Q I Rearranging the terms in equation (4-4.12) to group similar integrals gives 8A(W’ T(q)? q) = J K‘HV ' (k(rw‘”) + 9C“)? + V ' VIPC(r)vl - Zlej(z)6(r_dj)}e dr dz J: + GTE-(IL [q(r, t) — qpn-(r, t)]Aq(r, t) dr (1'! 3 + 2 13C, ‘ij j{qu(r, z) + [k(r)%%’+ pC(r)(v - fi)w]9} dr dz “F4 i=1 + JPCUXWG )|t=t dr (44.13) 9 j 82 Fix the Lagrange multiplier, w(r, t), as follows: J V . [k(r)Vw]+pC(r)%V +v . VIPC(r)\Vl— z e,(t)8(r—d,-> = n. j=1 (r) in 9 (44.14) (r0 < r s tf) -kz—a,-wtr,z>+Ihz—pC(r>1ww) = 0. (')°"F"(’=1’2’3) (44.15» 3" (t0(p")- — J(q —p p"): 22f1T—Y(z)1 dz j=l 1 ff ,2 _ _ n n 2 2 +20tT zOJnlq (r, t) qpn.(r,t) p p (r, t)] dr dz(4 5.11) Substituting the variation in the calculated temperature from equation (4-5.8) gives <1>dz 1:1 j—l +2 +12]: [T(dj’tzqn )-Y (1)] 2(111+ j=l l f n n n n n ’1 2 iaTfioIr4{[q (r,t)-CI,,,,-(r, 01-29 p (f,t)[q (r’t)—qpri(rzt)]+lp p 031)] }dr d! (46.13) 87 Taking the derivative with respect to the step size and solving for the optimal step size [3" , as shown in equation (4-5.7), gives J 2 £:[T(dj, t;qn) — Yj(t)]é(dj, t)dt+ arfiflnpnu, t)[qn(§, t) ‘quz'("z t)]dr dr —1 (Suzi- J ‘ 2 2 £36”?t)]2dt+arj‘::jr4[p"(r,t)] dr dt j= 1 (4-5.l4) Utilizing the relationship between the directional derivative and gradient in equation (4- 3.3) and the directional derivative in equation (4-3.7), the numerator of this function is J 2 £3“sz t;q") ‘ Yz(’)]é(djz ”‘1’ + arJZjBPWr, t)[q"(~£, t) — Chm-(r, z)]dr dz = j= 1 (V1 (61"), p">E(4-5.15> The step size in equation (4-5. 14) is rewritten as IL. VJ(r, t, q")p"(r, t)dr dt A ’0 4 p" : : J J 2 fléwj, t)]2dt + aTflfjr [pn(r, t)]2dr dt 2 "6911., ”"3; aTllpn”: i=1 0 " “ j=1 (VJ(q"), p">z: (4-5.16) 4-7.0 Implementation of the Gradient Method 4-7.1 Whole Domain The derivation to solve the IHCP using a gradient search method and adjoint equation approach have been presented. A summary of the solution with these methods, in the form of an algorithm for a whole domain implementation are given next. Equations presented in this section are identical to those derived earlier and are summarized here for clarity. 88 1) Select initial values for the function q°(r,t) (n = 0). A common choice is q0(r, t) = 0. If prior information is available set an-(r, t). 2 i) Solve the direct problem for the temperature T(r, t;q") V-(k(r)VT(r,t)) = pC(r)[—a—T(r,t)+v-VT(r,t)], (r) i" 52 (4-6.1) at (10 < t s. If) _k, inr, z)+h, T(r, z) = f2.(r, z), (r) on 13"“ =1’2’3) (4—6.2a) 3" , (tO <15 1f) -k(r)3.-T(r. t) = q"(r, t) . (r) on 1‘4 (4-6.2b) 3" (r0 < z g If) T(r, 0) = T0(r) , (r) in Q (4-6.2c) 2 ii) Evaluate the residuals eJ-(ti) = Yj(t,-) — T(dj, ti;q") (4-6.3) 3 i) Solve the adjoint problem to determine w(r, I) a J V . (k(r)V\1I(r, z)) + [pC(r)§—t-w(r, z) + v - V[pC(r)w(r, z))] + X ej(t)5(r—-dj) = 0 j: 1 (r) in Q (4-6.4) (r0 < t S If) -k.- 587w. z) + 1h.-— pC(r11v-zz)1w(r.z) = o, ”’ °" F" (' z" 2’ 3) (4-65a) n (10 < z g If) k(r) give, z) + pC(r)zs 2 (4-6 10) IIVJ(q"-1)Ilz: 4 i) Solve the (adapted) sensitivity problem for 8(5, t) V - (k(r)Vé(r, t)) = pC(5)[-a—é(r, t) + v ~ V60, 1)], (r) 1n Q (4-6.l 1) at (10 < t _<_ tf) (i =1, 2, 3) —k2. 23—2260, 1) + h,- é(r, t) = O, (r) on I",- (4-6.12a) (r0E p - J 2 2 _ J 2 2 f - I n - n .lejulewj, 1)] dt+a7£0jr4lp (r, 1)] dr dt _21I|9(d,-.t)llg+azllp HE J : j : (4-6.l3) 90 5) Compute the improved value for the heat flux q" + '(r, t) = q"(r, t) — fi"p"(r. t). (r) on F4 (MM) 6) Check for convergence of the estimated heat flux 2 ff],- [qnflvz t)-q"(r, 1)] 4’ dt = Ilq”+‘-q"||:~<€ (4-6.15) ‘0 4 If convergence has not been obtained let n = n + 1 and return to step 2 with the updated heat flux. If convergence has been met, use the residual principle to verify that the correct magnitude of (IT was used. The residual principle is J 15(q") = 2 J:’[T(d,-, q")—Y,-(t)]2a't=1:52 (4-6.16) 1:1 ° 2 . . . . . where 5 IS the expected no1se or error 1n the measurements and 1‘ 2 1 1s a relaxat1on parameter. An iterative regularization method can be obtained in the above algorithm by allowing the regularization parameter (IT to go zero. However, convergence is not evaluated by checking that by changes in the heat flux are small as in equation (4-6.15). Instead, in iter- ative regularization, the residual principal is used to select the number of iterations such that equation (4-6.l6) is satisfied and the sum-of-squares function is reduced to its expected value; once this is achieved the iteration process is stopped. Sequentially implementing this method is discussed next. Insight to the solution pro- cedures using the gradient method with adjoint equation approach is given in Section (4- 8.0) at the end of this chapter. 9 1 4-7 .2 Sequential Implementation The usual application of a gradient method with adjoint equation approach would solve the three problems over the whole time domain (tO < t S t f) , obtaining the estimated function q(r, t) for all time. Consequently, any inaccuracies in the estimated function at later times might influences the estimated function at early times. This dependence results from simultaneously estimating the function for all time. An alternative approach uses the sequential idea applied in function specification method. A sequential approach is quite effective for estimating the heat flux. In addition, it possesses other positive attributes. The components that are estimated at later times are not influenced by the early estimates, less computer storage (and possibly time) is required, and it is possible to linearize over the future interval and more efficiently solve non-linear problems. Benefits of a sequential method are discussed in more detail in Section 5-2.0. The sequential procedure can also be implemented for a gradient method using an adjoint equation approach. This procedure applies the method over smaller time intervals. Assuming that the function q(r, t) is known for times t S rm _ 1 , the unknown heat flux over r-future time steps, tm _ 1 < t S rm + r_1 is estimated. Equations for the whole domain implementation are identical to those for the sequential implementation; however, they are solved over the time interval tm _ 1 < t S tm + r _ 1 . After obtaining a solution, the time inter- val is shifted by one (or possibly more) time step(s) and the procedure is repeated until the whole time domain is covered. In contrast to function specification it is not required to assume a functional form for the heat flux, but such a procedure may be beneficial. With this sequential approach, many shorter duration problems are solved, compared to the whole domain approach that solves one large time duration problem. Operationally. it is 92 straight forward to use gradient search methods in a sequential manner. Implementing such a method, and the associated practical issues, are not as straight forward. Chapters 5 and 6 study and provide insight to this implementation. It is anticipated that this sequential approach will result in great reductions in the use of computer memory and have the potential to substantially reduce computational time. More savings are expected as the number of measurements become very large. These potential benefits are to be explored and are important parts of the proposed research. A more in-depth discussion of the benefits associated with the sequential approach is given in the subsequent chapters (Section 5-2.0). The formulation of the problem and the describing equations are nearly identical for a sequential implementation. However, the sequential implementation solves a series of smaller problems over the time domain. The time domain (to < t S t f) is broken into a series of sequential intervals each of length (tm_l < t S tm + r_ l) and solved for m = l, 2, . . . .M, where M = (tf- t0)/At— r, AI is the time step, and r is the number of future time steps. Assume that the heat flux for the sequential intervals up to and includ- ing time tm _ 1 is known. It is desired to compute the heat flux for subsequent times. Simi- lar to the whole domain implementation, the solution is presented in the form of an algorithm. 1 For the first sequential interval (m = 1) and iteration (n = 0) select the initial values for the function qO. Future sequential intervals (m > 1) use the estimates at the previous sequential interval to specify the initial value. A common choice is q0(r, t) = 0. If prior information is available set qpn-(r, t) . 93 2 i) Solve the direct problem for T(r, t;q") V - (k(r)VT(r, z)) = pC(r)[3T(r, z) +v . VT(r, n], (r) i“ Q (4-6.l7) at (tm_1 an (tin—lE 2 (4—6.26) IIVJ(q"“)||E 4 i) Solve the adapted sensitivity problem for 6(r, t) V (k(r)V6(r, t))= pC(r)|:(%- —6(r, t)+v V9(r, 0], (r) 1n 0 (4-6.27) (rm—l 15 (“529) z ||ézp,.Aq(r, z)dr dz 1' I F, k = 1: P M P M z + az 2 2 2 2 J0 Jr41¢.(z)q(r. z) — q,.,-t;) K which increases linearly to a magnitude of t;/2 then decreases linearly until reaching zero, and is zero elsewhere. The second variation is a step change in the heat flux 109 0 (t+<0) + + ‘12“ ) = i 1 (OSt+St;) (5-3.8) J O (t+>t;) which increases to a magnitude of one at t+ = 0 , remains constant until t+ = t; and then returns to zero. The final heat flux is a sinusoidal heat flux. It represents one-half the . 1t 31: . . perrod {—5, 3-) of a s1ne wave between zero and t+ = I; , and 1s zero elsewhere , O (t+<0) +z+ — 1 - ’+ 1 + + 539 ‘13()-< isrn—n2:+§ +1 ogzgzh ('-) ’zz [ O (t+>t;) A constant is added to the sinusoidal heat flux to maintain a nonzero flux. Analytical solutions are readily generated for these three cases. Solutions can be obtained using Greens functions (Beck et al., 1992). Although the test cases mentioned are standard cases and it is not difficult to solve for the temperature, closed form expressions are not typically given in the literature. Beck et al. (1985) give the closed form solution for a linear heat flux at the boundaries (x = 0, L). For completeness, the closed-form expres- sions for the temperature solutions with the three specified heat flux functions are pro- vided. For the triangular heat flux the solution is built using superposition of a heat flux which increases linearly with time (q+ = t+ ). The temperature due to the linearly varying heat flux is +2 +++_t 1+2+l 1M1+31+21 ¢l(x,t)——é—+[x —X '1' “It +[2—4x ~6x +6x —'4-5] +2 0° 1 2 2+ +7212 —4c cos(mrtx )[exp(——m 1: t )](5-3.10) The temperature solution for the triangular heat flux in equation (5-1.7) follows as u*<0) + + + + + .x , t <: 3< 41m". t“) — 2<1>I'(zc",t+ — III/2) (z;/2 < z+ _<. z;) 1 (q(x", 1+)— 2¢f(x*, z+ — z,j/2) + ¢f(x*, z” — z;) (,+ > ti) (5—3.l 1) Similarly, the temperature solution for the step help flux is built using superposition of a constant heat flux.The temperature solution for a constant heat flux (q+ = l ) is 00 ¢;(x+, t+)= t + [1x +2 —x+ + l] — —2— —1—2- cos(m1rx )exp(—m211:2 t ) (53.12) 2x 3 K2 17712 n = For the step change in the heat flux, equation (5-3.8), the temperature solution is (1+ <0) T+ + + — 4 + + + + + + + + + + + + \ ¢2(X 3t )—¢2(X rt _th) (t+>t;) The sinusoidal heat flux in equation (5-3.9) is the sum of a sine function and a constant, hence, superposition of these solutions is used. The sinusoidal part of the heat flux, q+ = sin [—1t(21+/t; + l/2)] , has a temperature solution 111 + 00 th t+ l 1 ¢;(x+, f) = ECOS[—n[2-+ + 2]] +4t; 2 [m4 4 +2 2:|cos(m11‘.x+) th m=1 “’1. +41: + + {21tcos[—1t[2-t: + a] + mzrtzt;sin [—1c(2t—+ + g] + mZTtZIZexp(—m21t2t+)} (S-3.14) th ’11 The temperature solution for the sinusoidal heat flux in equation (5-3.9) is 1 0 (t+ < 0) 73oz“, z“) = i1 130*, z“) + ¢§(x+. z”) (0 S ti 5 ’Z) 1 ¢§§(x+,t+ — 12) (z+ > t2) (5315) Data for the simulated temperature and heat flux are shown in Figure 5-1, 5-2, and 5- 3 for the triangular, step, and sinusoidal heat flux. All three cases demonstrate the effect that makes the IHCP difficult, mainly the lagged or delayed response in the temperature at the back surface due to the heat flux on the opposite surface. The temperature response is delayed relative to the application of the surface heat flux. The more gradual the variation in the surface heat flux is, the greater this delay becomes, as seen by comparing Figure (5- 2) and (5-3). The temperature response at the sensor location is damped also. That is, the magnitude of the temperature response at the sensor location is smaller that the response at the surface where the heat flux is estimated. These temperature data are used to estimate the surface heat flux. 112 0.6 I I F 0.5 - _ 0.4— q mm I 0.3 A 0.2 r 2 Temperature / Heat flux (dimensionless) Figure 5-1 Simulated temperature and heat flux data for the one-dimensional triangular heat flux 113 1.2 I I I 0.4 - I 0.2 Temperature / Heat flux (dimensionless) 1.5 2 Figure 5-2 Simulated temperature and heat flux data for the one-dimensional step heat flux test case 114 1 I I T I A 0.8»- m as .5 T*(1,:*) <0 _ + E 3 E? a: 0.4L 2;; Q) m \ 8 0.2“ D a H 0) Q. E o_ [— _O.2 1 l l L -0.5 0 0.5 1 1.5 Figure 5-3 Simulated temperature and heat flux data for the one-dimensional sinusoidal heat flux test case 115 5-4.0 1D Results - Simulated Measurements One-dimensional test cases are presented to study the sequential implementation. Implementing such a method is not recommended for the usual one-dimensional prob- lems; it is applied in this case to develop an understanding of this new method. (Nonlinear problems may be an exception for using the sequential method for one-dimensional prob- lems.) Three functional forms of the heat flux, triangular, step, and sinusoidal, serve as test cases. Investigations with exact data and data corrupted with errors are conducted. Errors, denoted 8n(dj, t) , are assumed additive Y(dj, t) = T(dj, t) + en(dj, t) (54.1) with a normal distribution, zero mean, constant variance, and uncorrelated. The true tem- perature T(dj, t) is computed from the analytical solutions, see Section 5-3.0. Random errors are generated with routines in MATLAB. 5-4.l Exact Data (no Measurement Errors) The three test cases are analyzed with exact data (an = O) for a dimensionless time step of 0.06 and a total of thirty time steps. Values selected for r are five, eight, and ten. Results are shown for the triangular heat flux in Figure 5-4. Graphic results are not shown for the step and sinusoidal test cases using exact data, but the analyses are discussed. Esti— mates for the final sequential interval (t f — rAt < t < If) at the end of the time region are not considered reliable and not reported. The results of the estimations are similar with less than a five-percent difference in the whole domain and sequential implementation; errors are largest near areas of sudden change in the surface heat flux. 116 Heat flux (dimensionless) 0.8 ' T T T I I t I . I o——oGM 4 H SGM r=5 0 6 _ <>"—">SGI\A r=8 .. ' H SGM r=10 — Exact 0.4 - - 0.2 - . 0.0 " i-T "3‘: ;;:: _ -02 . I 1 1 1 1 . 1 1 1 . -1.0 -0.5 0.0 0.5 1.0 1.5 2.0 + t Figure 5-4 Estimated heat flux for one-dimensional triangular test case with exact data 117 Table 5-1 Estimation results for exact simulated data Iterations . Tikhonov Comp S 1’ 0 D Analysrs per time 2 Method domain 0‘7" total seq int (sec) (°C) W/m Triangular Flux (O’ = 00°C) GM whole 0.001 7 - 2.14 8.59 13-04 0.0094 SGM r = 5 0.0001 1 78 3.0 3.64 8.64 E-04 0.0046 SGM r = 8 0.00087 87 3.8 6.86 8.72 E-04 0.0094 SGM r = 10 0.00107 87 4.1 8.64 8.72 E-04 0.011 Step Flux(6 = 00°C) GM whole 0.001 8 - 2.38 3.12 E-03 0.116 SGM r = 5 0.00014 79 3.0 3.56 3.17 13-03 0.109 SGM r = 8 0.0009 96 4.2 7.29 3.12 13-03 0.097 SGM r = 10 0.001 102 4.9 9.85 3.08 E—03 0.101 Sinusoidal Flux (0' = (10°C) GM whole 0.001 6 - 1.85 1.71 E-03 0.043 SGM r = 5 0.00012 80 3.1 3.48 1.70 13-03 0.044 SGM r = 8 0.0009 84 3.7 6.38 1.71 E-03 0.051 SGM r = 10 0.001 86 4| 8.34 1.71 E-03 0.050 Results from the analysis with exact data are given in Table 5-1 for estimating the three heat flux functions. The first column identifies the method, whole domain gradient method (GM) or a sequential gradient method (SGM). The second column lists the analy- sis domain investigated, either whole domain or sequential with the value of r specified. The Tikhonov regularization parameter is given in column three. Since exact data are used, the regularization parameter is specified as 0.001 for the whole domain analysis and 118 varied to produce the same sum-of—squares error (column seven) for the sequential analy- sis. The number of iterations, both total and average number per sequential interval, are given in columns four and five. Computational time is listed in column six. Measures of the error in temperature and heat flux are given in columns seven and eight, respectively. The error in the temperature is computed as 1 J N 5 l 2 S}, = {mg 2 [T(dj,t,.)—Yj(ti)] } (5-4.2) j=1i=l Equation (5—4.2) represents a measure of the error in the sum-of-squares function J S in equation (4-2.4). The residual (discrepancy) principle states that the error, S y in equation (5-4.2), should be reduced to its expected value, Alifanov (1994). Error in the estimated heat flux is 1 P N 5 A 1 A CD = {mkx z [(1070 t,-)-61(rk, 1,112} (5-4.3) = 11': 1 where q(rk, ti) is the true heat flux at location rk on F4 and time ti and 21 is the estimated value with no measurement errors. Equation (5-4.3) is referred to as the deterministic bias, or just bias, (Beck et al., 1985), since the estimated heat flux is computed without mea- surement errors. The deterministic bias represents the bias that is introduced by the inverse method itself; some bias is required to stabilize the ill-posed problem. The estimated heat flux can more closely approximate the triangular shape for small r-values using the SGM, particularly near the peak, see Figure 5—4. A smaller Tikhonov parameter permits the SGM to more accurately reproduce the peak, with less bias or smoothing of the estimated heat flux. It is interesting that the Tikhonov parameter is 119 reduced for the SGM at small values of r. The problem is more ill-posed for the SGM compared to GM. Hence, it was anticipated that the Tikhonov parameter would increase, but it decreases instead. The decrease in the Tikhonov parameter is required, even though the problem is more ill—posed, to permit the solution the “flexibility” to reduce the sum-of- squares to the magnitude specified by the residual principle. Regularization, tends to “stiffen” the inverse problem. When solved in a sequential manner, estimates near the end of the sequential interval are biased low (discussed below). To compensate for the biased estimates near the end of the interval the Tikhonov parameter must be reduced to permit “flexibility” in the first components and obtain the correct magnitude for the sum-of- squares. Note that as r is increased, the Tikhonov parameter also increases towards the magnitude specified for the whole domain solution. Several observations are drawn from the analysis with exact data. The computational time required for the sequential implementation was significantly more (by a factor or 2 or greater) than the whole domain solution. Larger magnitudes of r required additional com- putational time. Five was the minimum number of future times that could be used, which represents a dimensionless time based on the sensor depth of t: = 0.3. Fewer than five future steps, or more importantly a dimensionless time of 0.3, results in the estimated heat flux being biased low. The additional computational time required and minimum number of future time steps for the sequential implementation are explained by examining the characteristics of the gradient method near the end of the time interval. For illustrative purposes, consider that the steepest descent method is the iterative search method used. In this case the search direction is (the negative of) the gradient, which is related to the adjoint function. The heat flux at iteration n + 1 is 120 q" +1(ht) = q"(r, t) — 15571:, t). r 3 F4 (5-4.4) For the steepest descent method this becomes q“ 10‘, t) = qn(f, t)-(3"VJ(r, t),r3F4 (5-4.5) Substituting for the gradient from the relationship developed for the adjoint function, gives q" + 1(r, z) = q”(r, t) — p"{w(r, t) + aT[q"(r, t) — qpn-(r, 0]} , r 3 r4 (5-4.6) Recalling that the adjoint function is prescribed to be zero at the final time, equation (4- 6.21d), equation (5-4.6) shows that updates to the heat flux approach zero at the final time. Consequently, the estimated heat flux near the final time is biased. If no prior information is used, q pn-(r, t) = 0, the estimated heat flux at the final time is approximately (the dif- ference is the Tikhonov regularization term) the same as the initial guess (n = 0), q" + 1(r, tm + r__ 1) z q0(r, tm + r _ 1) . For a zero initial guess the estimated value at the final time is identically zero, q" + 1(r, t 0. m + r— 1) = It is not unreasonable to specify that the flux components near the final time approach zero. Nor is it unreasonable to specify that the adjoint function be zero at the final time. Very little information is available concerning the flux components near the final times. Due to the lagging effect of the temperature response at the senor location to variations in the surface heat flux, there is not time enough for information about the components near the final times to reach the sensor. The magnitude of the lagging effect depends on the dimensionless time (based on the sensor depth). For the same reason, the adjoint function, which is shown to be closely related to the gradient, is insensitive to the components near the end of the time interval. Appropriately, it is specified as zero there to simplify its com- putation. 121 The inherent difficulty of estimating the heat flux near the end of the time domain is more influential for the sequential implementation because a smaller number of time steps are considered. As stated previously, for sequential intervals shorter than a dimensionless time of 0.3 (five steps for these cases) results in biasing of the estimated heat flux. At the threshold of 0.3, the response at the sensor and sensitivity is large enough such that the estimated heat flux is not biased at the first time step of the sequential interval (the one retained for a sequential interval). Components beyond the first time step are biased due to the insensitivity of the temperature to these components. An anticipated benefit of a sequential implementation is an improvement in the com- putational efficiency. The expected improvement is based on the premise that after the first sequential interval (t0 < t < tO + rAt) an accurate initial guess for the estimated heat flux is available from the previous sequential interval. With an accurate initial guess it was expected that few iterations, possibly one, would be required to converge after the first sequential interval. Computational savings may not be realized for the one-dimensional problem because the conjugate gradient method is quite efficient. Results indicate that approximately the same number of iterations are required for all sequential intervals. A result that indicates the conjugate gradient search method is insensitive to the initial esti- mate. With a seemingly accurate initial estimate, convergence is expected within a couple of iterations, especially for the one-dimensional problem and should improve after the first sequential interval. This was not realized. the reason is discussed next. The computational time required for the whole domain approach was 50-90% less than the time required for a sequential implementation for r=5, see Table 5-1. As more future time steps were considered. the computational time increases significantly, nearly 122 doubling for an increase from r=5 to r=8. Additional iterations, which in turn increases the computational time, are required for the sequential implementation because of the biasing in the estimated heat flux near the end of the sequential time interval. The biasing of the estimates to zero (or a prescribed constant) results from the definition of the adjoint function, which is prescribed to be zero at the final time. The sequential implementation accentuates this effect, which was not expected to have such a significant effect. This was not expected to be significant because only data at the first time step is checked for the convergence. What was not realized was that in a sequential interval, the estimates at the end of the time interval influence the estimates at the beginning of the time interval. Fur- thermore, predicting the effect of the biasing is a difficult proposition. Its influence depends on the variation of the (unknown) heat flux. Such an effect is difficult to predict or anticipate. Examining the estimates at each iteration for a sequential interval gives insight to this effect. Estimated values for the triangular heat flux at each iteration for the sequential interval beginning at I: = 0.12 are shown in Figure S-Sa, b, and c, for r-values of 5, 8 and 10, respectively. For r=5 in Figure 5-5a the initial estimate (n = 0) , set from values at the previous sequential interval, is inaccurate. Biasing of the estimated heat flux resulted in a 40% error in this component during the previous sequential interval. At larger magnitudes of r in Figure 5-5b and c, the effect at the beginning of the sequential interval is less pro- nounced. However, the estimate at the first time step, which is the component retained for a sequential interval, varies with iteration. For r=10 the converged estimate (at n=4) for the first component varies only 2-3% compared to the initial estimate (n=0). The first com- ponent varies 10-30% while iterating (n=1 and 2). Variations in the estimates are caused 123 0.80 ' I I I I I ' I ' I ' I ' I G—ano 13—E1n=1 <>—<>n=2 3;, 0.60 r 1 “1:3 'QE’ — Exact .2 a 0.) .§. 3 0.40 — ~ é a: g 1 I 0.20 r . 0”8.00 I 0.10 L 0.20 ‘ 0.30 ‘040 I 0.50 l 0.60 I 0.701 0.80 + t Figure S-Sa Estimated one-dimensional triangular heat flux with no measurement errors for sequential interval beginning at time 0.12 for r = 5 124 0.80 i3 .5 § .§ 3 0.40 § Q: § I 0.20- 0“8.00 010 I 0.201 0.304 0.40 l 1 + t G—ano [El—El n=1 <>—<>n=2 A—An=3 ak—alenz4 — Exact . A; 0.50 0.70 L 0.80 Figure 5-5b Estimated one-dimensional triangular heat flux with no measurement errors for sequential interval beginning at time 0.12 for r = 8 125 0.80'1e111r1vrr 0.60 . E g 0.40 - § .5 1 3 ’5‘ E 0.20 ~ 8 I 4 J 4 G—9n=0 El—Eln=1 <>—<>n=2 A—An=3 l Hn=4 — Exact . 0.08 + t 0010101020 0.3 0.4010.5010.601.7010.80 Figure 5-5c Estimated one-dimensional triangular heat flux with no measurement errors for sequential interval beginning at time 0.12 for r = 10 126 by the biasing at the end of the time region influencing estimates throughout the entire time region. In particular, estimates four to five time steps from the end (within a dimen- sionless time of 0.3) have the largest change, and hence the most influence on the required number of iterations. These components, not the one at the end of the interval, are the dif- ficult ones. There is not enough sensitivity to estimate these component accurately, but there is enough to cause them to vary. Similar results were seen when estimating a con- stant heat flux. Estimating the heat flux with exact data has indicated that a sequential implementa- tion gives estimates within 5% (relative to maximum flux) of the whole domain method. Computational time required for the sequential solution was significantly greater. Compu- tational time is 50-90% greater than the whole domain, and increases depending on the magnitude of r. Additional computational time is required because the sequential imple- mentation is plagued by the inherent biasing of the components near the end of the time interval. This issue is addressed later by introducing some functional forms (function specification) over the sequential interval to control this effect. 5-4.2 Corrupted Data (Measurement Errors) An important issue when investigating methods for solving the IHCP is the effect of measurement errors on the solution. Methods that are typically insensitive to measurement errors do so at the cost of introducing bias in the estimates. Conversely, methods that are unbiased or have a small bias, are typically sensitive to measurement errors. These con- flicting objectives, minimal bias and sensitivity to measurement errors, are further dis- cussed later. The one-dimensional simulated test cases were evaluated again after adding 127 random measurement errors to the experimental data. Only errors in the measured temper- ature are investigated. Errors have standard statistical assumptions and are described in equation (5-4. 1). Estimation of the triangular, step, and sinusoidal heat flux with measurement errors added to the data are shown in Figure 5-6, 5-7, and 5-8, respectively. Results from the 0.8 v.vivr...,. 0——oGM B—BSGMrzs 06~ HSGMr=8 ‘ ' HSGM r=10 —Exact .5 0.4— . 2 <1) .§ 3 E? if 0.2~ - S I 0.0- - _0.2 1 1 . 1 L 1 L 1 1 1 L -1.0 -O.5 0.0 0.5 1.0 1.5 2.0 + t Figure 5-6 Estimated heat flux for one-dimensional triangular test case with measurement errors 128 1.5 ' Tl ' I ' I I I ' I e—eGM 9—9 SGM r=5 G—OSGM r=8 1.0 ~ A—ASGM r=10 ‘ :3 -— Exact 8 'E .2 :5; 0.5 ~ - é C1: 6 m 0.0 - . _05 . 1 L 1 . 1 L 1 L 1 L -0.6 -0.1 0.4 0.9 1.4 1.9 2.5 + t Figure 5-7 Estimated heat flux for one-dimensional step test case with measurement errors 129 Heat flux (dimensionless) 1.5 . . . , . 1 , r L 1 o—oGM H SGM r=5 HSGM r=8 1.0 ~ 111—1186M r=10 ‘ — Exact 0.5 - - 0.0 — - '0.5 1 1 1 1 L 1 1 1 1 1 1 -O.6 -O.1 0.4 0.9 1.4 1.9 2.5 + t Figure 5-8 Estimated heat flux for one-dimensional sinusoidal test case with measurement errors 130 Table 5-2 Estimation results for simulated data corrupted with errors Iterations ,1 Analysis Tikhonov per (I32? SY 05' Method domain (1.7 total seq int (sec) (°C) W/m2 Triangular Flux (0' = 0.0036°C ) GM whole 0.0028 6 - 1.84 3.63 13-03 0.034 SGM r = 5 0.0021 77 2.9 3.48 3.65 E—03 0.050 SGM r = 8 0.0019 91 4.0 6.95 3.64 E-03 0.040 SGM r = 10 0.0026 88 4.2 8.50 3.62 E-03 0.039 Step Flux (0' = 0.012°C) GM whole 0.0033 7 - 2.06 1.23 E-02 0.139 SGM r = 5 0.0021 76 2.9 3.43 1.25 E-02 0.147 SGM r = 8 0.0026 87 3.8 6.66 1.24 E-02 0.1 19 SGM r = 10 0.0030 82 3.9 7.74 1.23 13-02 0.101 Sinusoidal Flux ( 0' = 0.006°C ) GM whole 0.0019 7 - 2.15 6.19 13-03 0.061 SGM r = 5 0.0016 76 2.9 3.42 6.19 13-03 0.088 SGM r = 8 0.0015 91 4.0 6.91 6.19 13-03 0.071 SGM r = 10 0.0019 92 4.4 8.88 6.19 13-03 0.070 analysis are given in Table 5-2. The tabulated results are similar to that presented for the analysis with exact data. The type of analysis, whole domain (GM) or sequential (SGM), Tikhonov parameter, number of iterations, computational time, and temperature and heat flux errors are given in the table. In this case, which has measurement errors, the error in the estimated heat flux represents an error due to bias in the algorithm and the influence of measurement errors. This error is defined as the mean-squared error. The computation of 131 this error is identical to computation of the previous error in heat flux; however, the heat flux is estimated with random errors present in the measurements 1 P N 2 A 1 A GS, : {P(N _1) 2 2140700)" qe(rk2 ti)]2} (5-4.7) k = 11': 1 where 218 is the estimated heat flux with measurement errors in the temperature. The residual principle was used to select the magnitude of the Tikhonov regulariza- tion parameter. As stated in the principle, the sum-of-squares function should be reduced to the expected level. Since random errors are used, this is specified as the standard devia- tion of the measurement errors. The analysis with measurement errors, as expected, demonstrates similar computa- tional aspects as the analysis without measurement errors, requiring significantly more time for the sequential method. It also showed that the whole domain method was superior to the sequential implementation in the accuracy of the estimated heat flux. Errors in the estimated flux were up to 50% less, depending on the magnitude of r, and were largest for smaller magnitudes of r. (Error in the estimated step heat flux was less for r—values of eight and ten; however, this did not include the step down in the heat flux because this region was within r steps of the final time.) These results show that, similar to the function speci- fication method, the sequential gradient method requires a pr0per selection of r. However, it further shows that selecting a magnitude of r and the Tikhonov parameter (r, (1T) , such that the residual principle is met, does not insure optimum results. Improved results may be obtained with another set of parameters. For smaller magnitudes of r the sequential method is more sensitive to measurement errors, but the method reduces (IT to meet the 132 residual principle. Although the whole domain method provides a more accurate estimate, the sequential method is competitive for a proper selection of r. If r is made large enough, the estimates are independent of r and are the same for both the sequential and whole domain solutions. Quantifying the results of an estimation are discussed in Beck et al., (1985) and Scott and Beck (1989). The errors suggested for quantifying the results of an estimation are the deterministic error D and the mean-squared errorSe. These errors provide insight into the characteristics of an estimation algorithm. The errors are computed in equation (5-4.3) and (5-4.7) and are tabulated in Tables 5-1 and 5-2. For the deterministic error the heat flux is estimated with no measurement errors, while for the mean-squared error the heat flux is estimated with random errors in the measurements. The deterministic errors, which show the bias of an algorithm, are smaller for the sequential approach for the r=5. The deterministic bias, in the final column of Table 5-1, is smaller because OtT is smaller for all three cases for r=5. Since there is less regularization the solution can more quickly respond to changes in the heat flux. However, as r increases (and a7 increases) the deterministic bias for the SGM becomes larger, and equals or exceeds the whole domain value. The mean-squared error, which can be compared to the deterministic error to determine the sensitivity to measurement errors, is exclusively better for a whole domain solution. The mean-squared error is shown in the final column of Table 5-2. A sequential method permits the Tikhonov parameter to be smaller at low r—val- ues, however, this is at the cost of an increased sensitivity to measurement errors, as shown by a larger mean-squared error at these r—values. As r increases the SGM becomes less sensitive to measurement errors, of course in the limit the SGM and GM are the same. 133 Although the SGM provides comparable results to the GM, the computational requirements are significantly greater. A 100% (or more) increase in the computational time is typical for a sequential implementation compared to a whole domain solution. The increase in computational time is a direct consequence of the sequential implementation (of length r—steps) requiring approximately half as many iterations for each sequential interval as the whole domain requires for a complete solution. Since these results are for the one—dimensional problem, when the methods are extended to two-dimensions the requirements for the sequential implementation might be greater in comparison to the whole domain approach. As demonstrated previously, the additional computational requirements are caused by the inability to estimate components near the end of the sequential interval. Several approach were investigated to improve the computational requirements of the sequential implementation. It was clear from observing the results during a sequential interval (see Figure 5-5) that the variation in components near the end of the interval result in an increased number of iterations. If these components over the end of the sequential interval could be controlled, or anticipated in their response, improvement in computa- tional requirements is possible. Note that the components in the range 4-5 steps from the end of the interval (a dimensionless time of 0.25 to 0.3) are troublesome. On either side of these components the response of the heat flux is understood. Approaches to improve the computational efficiency of the sequential method looked at three aspects of the sequential implementation. First is the setting of the initial condition for a sequential interval. Next, was using regularization to control the variation of certain components in a sequential interval. Lastly, was the use of prescribed functions over the 134 sequential interval, similar to the function specification method. As it turned out, only the final approach, using function specification over the sequential interval significantly improved the computational efficiency. Much was learned concerning the method, how- ever, and is described below. Setting the initial condition for the subsequent sequential interval, from the previous interval, has all the required information except for I-components, where I is the number of components retained on time for a sequential interval. Initially, all values from the past interval were carried to the next interval then the final I-values were specified to be zero. But it was noticed that the values near the end of the interval did not change with sequen- tial interval. Another procedure was used. Estimates from the previous interval were used from the beginning and end of the sequential interval. The procedure is shown graphically in Figure 5-9. For a specified number of time steps, m end (typically m en d = 4) from the end of the interval, the initial values were set from the previous interval 0 q (r’t) : qNC0"V(f,I—IAI) ’tm+r-1—mmd+lStStm-+r—l (5-4'8) where qum’ is the heat flux estimated at the previous sequential interval and q0 is the initial estimate for the next interval. Equation (5-4.8) specifies that the initial values at the end of the sequential interval should be the same as they were at the end of the previous sequential interval. The components are shifted I-values in time because the time interval advances l—values for the next sequential interval. Initial values at the beginning of the time interval are specified directly from the previous interval 0 q (r, I) = qNCOflV(’-,t) ’tm—1+IStStm+r-l—m (5-4.9) rnd 135 1! tm+r—l+l N / I. N / I / + 1. // / E _1 / j / / g 1. / / l’ +/ / / E / s ‘7 / 1 :,/ .l +' ‘ I E L f “ '- /’ 1 .5: 1 / I '/ T a . /' .3 // .1 o E , ..1 E // - 33 l 1 _ / ---- 45 | / ' t: 1- " .1 ' o +_ + . j‘ S1 E 1 ,2 N 1 E” o 3 '1 1 U \1 . .... I E . « l _______ #L 1'. “N ‘~ + + 5 NE N --~ I -1 :_ -+ H)- + —————— h;- “E E H ‘ -1~ E ,... ‘ U" (1 'SIMIOJNb Figure 5-9 Procedure for specifying the initial condition using converged estimates from the previous sequential interval 136 This leaves I-values (where I is the number of retained components on time) to be speci- fied, which are computed using linear interpolation to join the two previously defined regions. Unfortunately, accurately specifying the initial condition from the previous interval was not causing the increased computations. Computational requirements were unchanged using the procedure outlined for setting the initial estimated. Rather, it was understanding that the estimated heat flux in the previous sequential interval was not as accurate as antic- ipated, particularly components about 4-5 steps (or t: = 0.3) from the end of a sequential interval. These components vary significantly from one sequential interval to the next, which influences other components in the interval. The results in Figure S-Sa, b, and c show that the initial condition was not as accurate as believed. Furthermore, the components that cause the most difficulty in the sequential estimation are the last 4-5 components (covering the final 0.3 in dimensionless time of the interval). Consequently, regularization was used to attempt to control the change in these components. The Tikhonov regularization parameter was specified to be a function of time, with magnitudes larger near the end of the sequential interval. The larger regulariza- tion parameter drives the estimated heat flux to zero. This approach did eliminate changes in the component near the end of the sequential. But, the changes that would have occurred in these component migrated to other components. In retrospect this should have been anticipated, since the solution for the heat flux should conserve the total energy (for a specified r-value), by forcing some components to be zero (or a constant), other compo- nents will accommodate to conserve the total energy. Computational aspects were not improved. 137 A positive aspect of the gradient methods is their function estimation basis. The diffi- culties encountered in the sequential implementation, however, are due to the attempt to estimate a function over the sequential interval, even though, only one component of the function is typically retained. A final approach to reduce the computational requirements is to constrain the heat flux over the sequential interval so that not as much computational effort is needed to estimated all components on a sequential interval. Fewer parameters are estimated instead of a function. Specifying several functional forms over the sequential interval were tried: cubic, parabolic, linear, and constant. Additional constraints were imposed for the higher order functions, such as zero slope or zero second derivative at the end of the sequential interval, as required. Prescribing a single function over the sequential interval did not improve the computational aspects when higher order functions were used. However, specifying the heat flux be constant over the sequential interval did improve computational efficiency significantly. The triangular heat flux case was re-evaluated using function specification over the sequential interval. Herein, this method is referred to as the sequential function specifica- tion gradient method (SFSGM). Maintaining the heat flux constant over the sequential interval stabilizes the problem and the Tikhonov regularization parameter is set to zero. The results for the exact data are shown in Table 5-3 for the triangular heat flux using the function specification assumption; graphical results are not shown. The accuracy of the estimated heat flux is comparable to the whole domain. A direct comparison is not made because the Residual principle is not met by all r-values. The magnitude of the sum-of- squares error suggest an r between 4 and 5. In this range the error in the estimated heat flux is comparable, and more importantly computational requirements are reduced. For 138 Table 5-3 Estimation results for prescribing a constant heat flux over the sequential interval Analysis Tikhonov 1“”er C113: SY GD 0" 63' Method domain aT total seq int (sec) ( o C) W/m 2 Triangular Flux (G = 00°C) GM whole 0.001 7 - 2.14 8.59 E-04 0.0094 SFSGM r = 3 0.0 51 1.8 1.08 9.65 E-05 0.0054 SFSGM = 4 0.0 49 1.8 1.47 6.07 E-04 0.0052 SFSGM r = 5 0.0 48 1.9 1.90 1.89 E-03 0.014 SFSGM r = 6 0.0 48 1.9 2.24 9.13 E-04 0.025 Triangular Flux (0' = 0.0036°C) GM whole 0.0028 6 - 1.84 3.63 E-03 0.034 SFSGM = 4 0.0 54 2.0 1.62 2.52 E-03 0.096 SFSGM r = 5 0.0 50 1.9 1.98 3.27 E-03 0.039 SFSGM r = 6 0.0 50 2.0 2.38 4.99 E-03 0.040 SFSGM r = 8 0.0 46 2.0 2.93 1.07 E-02 0.048 r=5, using the function specification requirement (SFSGM) the computational time is reduced by nearly 50% compared to the standard SGM; computational time is 10% less than a whole domain solution. The heat flux estimated with measurement errors added to the data is shown in Figure 5-10 when assuming the heat flux is constant over the sequential interval. Estimation results are also given in Table 5-3. With the heat flux held constant over the sequential interval the sequential implementation improves significantly in computational efficiency. Since only one components is estimated at each sequential interval, convergence is 139 obtained with few iterations. When the heat flux is held constant over the sequential inter- val, denoted SFSGM, a sequential solution is competitive with the whole domain solution. 0.8 ' I ' l u T ' I 1 I r o—oGM ‘ H SFSGM r=4 0 6 _ " <+—<>SFSGM r=5 ”a? ' A——ASFSGM r=6 é’ HSFSGM r=8 .g — Exact t: E g 0 4 b ‘ >< =2 5 m 0.2 - - I}; .\ 0.0 ” L‘3'16'.5; : 3 - ‘. _02 1 1 1 1 1 1 1 1 L 1 1 -1.0 -0.5 0.0 0.5 1.0 1.5 2. + t Figure 5-10 Estimated heat flux for one-dimensional triangular case with the heat flux held constant over future sequential interval 140 The computational time is further reduced for a sequential solution by using prior information. The prior information enters the solution in the Tikhonov regularization term of equation (4-2.4). Specifying the prior information as the initial estimate, which is set from values at the previous sequential interval, is shown to significantly improve the accu- racy of the estimated heat flux as well as reduce the computational time required for a sequential solution. The use of the prior information is discussed in the next chapter. Results using prior information are not shown for the one-dimensional case. 5-4.3 Results - Experimental Measurements Although simulated measurements provide insight to the behavior of the method, the true test is with experimental data. Gradient and iterative regularization methods are applied extensively in the literature to solve the IHCP. However, few of these investiga- tions have applied the method to experimental data. Measured temperature and heat flux histories from the experiment to estimate the thermal conductivity are used. For this pur- pose, the problem is reversed and the estimated thermal properties and measured tempera- tures are used to recover the transient surface heat flux. Since electric heaters were used to generate the heating, the heat flux is also quantified. Consequently, a comparison between measured and estimated heat flux can be made. The experimental apparatus is shown in Figure 2-1 and discussed in section 2-2.0. Electric heaters symmetrically heat the composite material while thermocouples measure the transient temperature. Since the experiment is symmetric, the measured power to the heaters quantifies the energy and heat flux to the specimens. In the thermal model of the experiment, shown in Figure 2-2, the thermal effects of the mica heater, and more 141 importantly the contact resistance, are included. An effective thermal conductivity, which represents the heater, contact resistance, and the cement used to install the thermocouples, is used to group these effects. Heat losses to the insulation are also included. Temperature measurements are made very close to the surface where the heat flux is estimated (a depth of 0.44 mm). However, the relatively low effective thermal conductivity of the mica heater/contact resistance results in the problem being more difficult than it appears. The Fourier number based on the depth of the closest sensor and thermal properties of the mica heater/contact resistance is approximately 0.06. The relative high thermal conductivity of the carbon-carbon specimen increases the difficulty of estimating the surface heat flux also. This is demonstrated by the limiting case, if the conductivity of the specimen were infinite, the sensitivity to the surface heat flux is zero and it is impossible to estimate it. This case was also investigated by Beck et al. (1996), which among other things investigated the use of the residual principle for selecting the number of future time steps for the function specification method. Since seven reliable thermocouples were used to measure the temperature at nominally the same location, the error in the measurements could be quantified. Beck et al. suggested using the estimated standard deviation of seven measurements from the average temperature at each time. The standard deviation was approximately 004°C at the beginning and end of the experiment when the heater was not active and 012°C when the heat flux was active. A time average of this standard devi- ation was 0.083 C. This value was used as the level of noise for selecting the regularization parameters with the residual principle. Beck et al. (1996) also compare several inverse methods. Function specification, Tikhonov regularization, iterative regularization, and specified functions over large time 142 regions using Green’s functions were compared. Results indicate that estimates from the various methods are quite comparable. Although computational requirements were not the focus of the study in Beck et al. (1996), it was noted that certain methods require consider- ably more computational time. Estimations of the surface heat flux are shown in Figure 5-11 for the experimental case. Several variations of the gradient method are compared to estimate the surface heat flux. Whole domain gradient method and sequential gradient method including specifying a constant flux over sequential internal are shown. A remarkable outcome of the different techniques is that about the same results is obtained. Results of the analysis are also listed in Table 54. Only for r=4 is there a significant difference between the whole domain and sequential solutions. Otherwise, the sequential solutions are as good as, or better than, a whole domain solution. The sequential solutions, denoted SGM in column one of Table 5- 4 require about three iterations per sequential interval. Holding the heat flux constant over the sequential interval, denoted SFSGM in column one, requires about two iterations per sequential interval. This represents a reduction of approximately 30% in computations with SFSGM. A larger decrease in the computational time is achieved because the number of iterations required for SFSGM is actually less. Although the second iteration is started, frequently the gradient computed during the second iteration is very small (less than 1E- 15) and based on the magnitude of the gradient iteration is stopped. So a full iteration is not completed (the sensitivity problem is not solved) and greater savings are realized. These additional savings increase with r. 143 9000.0 1 1 1 1 1 o—oGM B—EISGM r=8 — Measured A5000.0 - - ”E 1 E 6 {13 3000.0 - , - 8 I J 1000.0 - . -1000.0 1 ‘ 1 1 ' ' 0.0 10.0 20.0 30.0 40.0 time (seconds) Figure 5-11 Estimate heat flux for one-dimensional experimental case 1010#30.l 144 Table 5-4 Estimation results for analysis of experimental case Iterations Analysis Tikhonov per (£1912? SY 64 Method domain a7" total seq int (sec) (0 C) W/m2 Experiment Case 1010#30.1 GM whole 1.0 E-08 8 - 24.1 0.080 766 SGM r=4 4.0 E- 10 603 3.1 30.0 0.079 1400 SGM r=6 2.0 E-09 564 2.9 47.4 0.077 754 SGM r=8 4.5 E-09 560 2.9 64.7 0.079 716 SGM r=10 7.0 E-09 565 2.9 80.2 0.078 726 SFSGM 124 0.0 382 1.9 16.5 0.039 2250 SFSGM r=6 0.0 373 1.9 26.4 0.044 762 SFSGM r=8 0.0 347 1.8 33.1 0.057 699 SFSGM r=10 0.0 320 1.7 39.4 0.083 751 SFSGM r=6 3.5 E-lO 370 1.9 25.4 0.086 764 SFSGM m8 4.5 E-10 342 1.8 32.7 0.077 714 Experimental Case 1023 #4031 GM whole 4.0 E-09 8 - 25.3 0.080 1230 SGM i=6 9.0 E- 10 578 3.0 47.0 0.084 1182 SGM :8 2.0 E-09 578 3.0 82.1 0.084 1141 SGM :10 3.0 E-09 576 3.0 82.5 0.079 1167 SFSGM r=6 0.0 359 1.9 26.0 0.036 1234 SFSGM r=8 0.0 342 1.8 34.2 0.052 1162 SFSGM r=10 0.0 306 1.8 39.8 0.081 1261 SFSGM r=6 1.5 E- 10 359 1.9 26.0 0.080 1215 SFSGM r=8 2.0 E- 10 336 1.8 32.6 0.077 1180 SFSGM r=10 1.2 E- 10 307 1.6 39.9 0.084 1267 Chapter 6 APPLICATION OF THE SEQUENTIAL GRADIENT METHOD: TWO-DIMENSIONAL IHCP 6-1.0 Introduction The solution of the two-dimensional IHCP is investigated in this chapter. Use of the sequential gradient method (SGM), possibly coupled with prescribing a functional form over the sequential interval (SFSGM) are studied. A comparison of sequential gradient methods with the standard whole domain gradient method (GM) is made. Some compari- son with the standard function specification method, which does not use a gradient search method nor an adjoint equation approach, is presented. The two (or three) dimensional problem is the intended use of the proposed sequential methods. As demonstrated with the one-dimensional cases, assuming the heat flux is constant over the sequential interval improves the computational time of the sequential implementation, and makes it competi- tive with the whole domain approach. Additional steps may be required to improve the computational aspects for the two-dimensional problem also. An approach similar to the one-dimensional analysis is pursued for the two-dimen- sional case. Several test cases for which the surface heat flux is known are studied to quan- tify the accuracy of the method for a two-dimensional analysis. The two-dimensional cases will demonstrate the ability to estimate spatially dependent fluxes. In addition to 145 146 simulated cases, an experimental case with a measured surface heat flux, which is two- dimensional and transient, is analyzed. Several variations of the sequential gradient method are suggested in this chapter. Three methods are investigated. The standard whole domain gradient method (GM) solves for the heat flux with data from all measurement times. The sequential gradient method (SGM) solves for the heat flux using data from a sequential interval. The sequential func- tion specification gradient method (SFSGM) solves for the heat flux over a sequential interval while prescribing a functional form for the heat flux on time, but allows for spatial variation. In addition, the use of prior information is suggested during the sequential solu- tion to improve the sequential method. This is the first known application of prior informa- tion to improve a sequential solution. The remainder of this chapter is summarized. The calculation of the simulated tem- perature data is discussed in the next section. Section 6-3.1 presents results for the simu- lated test cases with exact data. Test cases with measurement errors added to the simulated data are discussed in Section 6-3.2. An experimental case, with a measured surface heat flux is analyzed in Section 6-4.0. 6-2.0 Simulated Temperature Data A rectangular domain of dimensions [0, a] x [0, b] is studied for the two-dimen- sional simulated cases. A schematic of the geometry is shown in Figure 6-1. One bound- ing surface of the two-dimensional body is assumed to have an unknown heat flux while all other surfaces are insulated. Furthermore, constant anisotropic thermal properties and a uniform initial temperature are assumed. 147 q(x1t) /‘ \ ’ 4 O O O O O O O O Jk T o / \ o 11 6 § 11 b 5 x h / \\\\\\ \\\\\\ \\\\\\ \\\\\\\\\\\\ \\\\\\ \\\\\ .\ I q = 0 y f a $» Figure 6-1 Two-dimensional geometry for simulated test cases The problem can be mathematically written as 2 2 [bzkaa T: a 7;" 6T?“ + — _ __ _ (6-2.10) aZky ax+2 ay+2 81+ aT?‘ _ a i = (13,114", 1*) = ? (6-2.lla) .v ),.:0 8T: = 0 (6-2.llb) + 3y 148 arf ; = 0 (6—2.l 1c) x x+ = 0 Br: I x+ = 1 T:(x+, y+, 0) = O (6-2.lle) where the following dimensionless variables were introduced q: = {-15}; (62.12) T-—T + l 0 = —— (6-2.l3) ' (qNb/k),) k,/ c)! + = ( " z (6-2.l4) b x+ = (1: (6-2.15) y+ = 11) (6-2.16) The subscript 1' indicates the prescribed heat flux. The estimation of two different pre- scribed heat fluxes is investigated for the two—dimensional case. A step change in heat flux on time (starting at t+ = 0 and ending t+ = t2) and position (from x+ = 0 to x)r = a1) is the first flux (1' = l) examined qi(x+,t*)= l (O 1 8. 5 A 2 05 o 0 0 0 1 Ammo—coficoEé 5E “mom oomtzm Figure 6-2 Simulated temperature and surface heat flux for a two-dimensional triangular heat flux 151 0.2 :2 :g : , \LLLL_____:% :52??? £22 55:2 , :3: :, —5 / / 7; 0.3 \ 0.2 \ Ammo—coicoEEv 11:58an .H. Ammo—coicoEEV xsc Hao: oomtsm Figure 6-3 Simulated temperature and surface heat flux for a two-dimensional step heat flux 152 depth is At: = 0.06 , which represents a difficult case; two-hundred time steps are consid- ered in the analysis. “Exact” simulated temperature data are generated from the finite ele- ment solution using TOPAZZD (Shapiro, 1986) The geometry shown in Figure 6-1 is discretized for the numerical solution. A finite control volume mesh with eleven nodes in the x-direction and twenty-one nodes in the y- direction is used. A more coarse node spacing is specified for the x-direction. Because eleven sensors are located along this surface, eleven components of heat flux are to be esti- mated. It is possible to define more nodes along this surface and estimate a greater number of spatial components (discussed later). Altemately, more nodes can be defined along the surface and fewer spatial components, compared to the number of nodes, can be estimated by introducing some spatial smoothing using basis functions; see Section 4—8.0. In this study, it is not desired to impose spatial smoothing on the solution. Instead the number of nodes defined for the numerical solution are set based on the number of available sensors. Hence, eleven nodes are defined on the surface. The number of nodes is increased and more spatial components are estimated in a later analysis. Care is exercised to not have too few of nodes and incur numerical error from a coarse discretization. The estimated surface heat flux for a whole domain solution (GM) and the sequential gradient method (SGM) with r=6 are shown in Figure 6-4a and b, respectively, for the tri- angular test case. The GM has the whole domain characteristic of over—shooting near regions where the heat flux has an abrupt change, at the beginning and ending times. The SGM reduces the effect of these characteristics. Both methods accurately estimate the sur- face heat flux. Heat flux (dimensionless) 153 a) GM Whole Domain 0.8 4 0.6 \1 0.4 x 0.2 \ 15 0.8 Figure 6-4a Estimated surface two-dimensional heat flux for triangular test case using data without measurement errors. Whole domain solution Heat flux (dimensionless) 154 b) SGM r = 6 Figure 6-4b Estimated surface two-dimensional heat flux for triangular test case using data without measurement errors. Sequential solution with r=6 155 Table 6-1 Estimation results for the two-dimensional IHCP with exact simulated data Analysis domain Iterations Tikhonov CPmP S Y 6 D per t1me Method 1' I “T total seq int (sec) (°C) W/m2 Triangular Flux (11 Components) GM all all 1.00 E04 1 l - 35.7 2.53 E-04 7.46 E-02 SGM 6 1 1.02 E-04 758 4.0 71.8 2.50 E-04 8.05 E-02 SGM 10 1 3.20 E04 735 4.0 121.2 2.52 E-04 7.79 E-02 SFSGM 6 l 1.05 E-05 244 1.3 23.2 2.51 E-04 7.89 E-02 SFSGM 10 l 0.0 238 1.3 35.6 4.29 E-04 7.32 E-02 Step Flux (11 Components) GM w all 1.0 E-04 12 - 38.8 4.42 E-04 9.30 E-02 SGM 1 1.25 E-04 762 4.0 71.7 4.46 E-04 9.92 E-02 SGM 10 1 4.05 E-04 737 4.0 119.9 4.43 E-04 9.54 E-02 SFSGM 6 1 1.67 E-05 264 1.4 24.9 4.42 E-04 9.68 E-02 SFSGM 10 l 0.0 260 1.4 42.1 5.41 E-04 8.84 E-02 Triangular Flux (21 Components) GM w all 1.00 E-04 11 - 66.3 2.53 E-04 7.46 E-02 SGM 6 1 1.01 E-04 758 3.9 130.7 2.50 E-04 8.12 E-02 SGM 10 1 3.10 E04 727 3.9 222.4 2.50 E04 7.72 E02 SFSGM 6 l 1.05 E-05 270 1.4 47.0 2.52 E-04 9.65 E-02 SFSGM 10 1 3.00 E-OS 278 1.5 83.6 4.91 E-04 7.89 E-02 Pertinent information about the estimations with exact data are given in Table 6-1. The table lists the solution method and analysis domain, with the number of future time steps and number of components retained on time in the first three columns. Column four gives the magnitude of the Tikhonov parameter. The number of iterations, both total and 156 average per sequential interval are given in columns five and six. Computational time is listed in column seven. The errors in the sum-of-squares temperature and estimated heat flux (bias error) are shown in columns eight and nine. Values of r = 6 and r = 10 were studied for the number of future time steps. The Tikhonov regularization parameter was specified as 0.0001 for the whole domain solution (GM). Subsequent sequential solutions (SGM and SFSGM) varied the Tikhonov parameter to match the magnitude of the sum-of- squares, S y, to the level of the whole domain solution. Estimation results with exact data for the triangular and step heat flux, given in Table 6-1, have trends similar to the one-dimensional case. A comparable heat flux to the whole domain (GM) solution is estimated with the sequential implementation (SGM), the bias error (last column of Table 6—1) is 6-8% larger for r = 6, but it requires approximately twice as much computational time. Increasing the r-value to ten reduces the bias error to a magnitude 2-4% larger than the GM, but requires a proportional increase in the computa- tional time, i.e. computational time increases by 167%(10/6) , compared to the computa- tional time for r = 6. In the one—dimensional solution the computational aspects were improved by main- taining the heat flux constant (on time) over the future sequential interval, which is the standard function specification assumption. The same procedure was tried for the two- dimensional problem. Prescribing that the heat flux remain constant over the sequential interval (method SFSGM in Table 6-1), requires one-third the computational time of the SGM for r = 6, and has a bias error that is only 4-6% larger than the whole domain esti- mation. Improved accuracy was obtained when r was increased to ten, bias error was 2-5% 157 less than whole domain solution. Notice that the Tikhonov parameter was set to zero for the SFSGM with r = 10. Even with this parameter set to zero the sum-of-squares was not reduced to the required level. This is due to the effects of the Tikhonov parameter and the function specification approximation, which are discussed next. The SFSGM has superior computational aspects. It requires 35% less computational time than the GM for both the triangular and step test case for r = 6; computational requirements are comparable for r = 10. One observes in Table 6-1 that the estimation with exact data requires a larger magni- tude for the Tikhonov parameter when solving sequentially compared to whole domain. Since a larger magnitude of the Tikhonov parameter is used, the sequential solution is more biased than the whole domain solution, particularly near regions of variation in the heat flux. An increase in the Tikhonov parameter is expected because the shorter duration sequential problem is more ill-posed than the whole domain problem. For the one-dimen- sional solution, in general, a smaller Tikhonov parameter was required for the sequential solution compared to the whole domain solution, contrary to what was expected. It is believed that the one-dimensional problem required a smaller Tikhonov parameter to per- mit the solution algorithm the flexibility to reduce the sum-of—squares magnitude to the desired level. The reduction in the Tikhonov parameter is greater at small r-values. As r is increased the Tikhonov parameter for SGM approaches the magnitude of GM, in the one- dimensional case. Applying the function specification approximation in the inverse solution provides significant regularization, which increases with r. As noted in Chapter 5, regularization imposes “stiffness” to the inverse solution. When maintaining the heat flux constant over 158 the sequential interval, considerable “stiffness” is imposed on the solution. Consequently, the Tikhonov parameter must be reduced in magnitude to permit obtaining the desired sum-of-squares magnitude, S y. At some level of r, the bias error becomes large enough such that even when the Tikhonov parameter is specified as zero, the sum-of-squares func- tion can not be reduced any further. This is a result of the bias introduced when assuming the heat flux is constant over the sequential interval. A noted benefit of the gradient method is that it is not required to specify the func- tional form of the unknown heat flux. Since a numerical solution is used, however, some approximation of the heat flux is required. The simplest numerical approximation of the heat flux is used, which assumes the heat flux is constant over a control surface surround- ing the node. Consequently, the functional heat flux is numerically represented by P spa- tial components, where P is the number of nodes along the surface of the unknown heat flux. Specifying more nodes along this surface will improve the numerical accuracy of the direct problem, but requires estimating more spatial components, which may increase the difficulty of the inverse problem depending on sensor locations. A general rule is that the number of spatial components estimated should not exceed the number of sensors. Fewer spatial components of the heat flux, compared to the number nodes, can be estimated by prescribing a spatial functional form using basis functions, see Section 4-8.0. Prescribing a spatial functional form introduces regularization and helps stabilize the solution. In the previous simulated cases, given in the first two blocks of Table 6-1, eleven nodes were defined along the x-direction and eleven spatial components were estimated. Now the triangular test case is analyzed with twenty-one nodes components on the surface and twenty-one spatial components are estimated, with eleven equally spaced sensors 159 along the surface. Results are given in the final block of Table 6-1. The estimated heat flux is nominally the same for the whole domain (GM); of course computational time is approximately doubled with twice as many nodes. The sequential method does not per- form as well. Bias error is increased by 1% for the SGM, but increases 22% for the SFSGM with r=6. When estimating more spatial components than available sensors with a sequential method the spatial components between the sensor locations are biased low for r=6. This produces spatial oscillations in the estimated heat flux, a “zig-zag” pattern about the true flux value exists in the estimated heat flux. The oscillations are reduced by increasing the magnitude of r. It appears that estimating more spatial components than available sensors is possible using the gradient method. At least for the case when sensors are evenly distributed along the surface it is possible. The method seems to have some spatial regularization inherent. Implementing a sequential solution degrades this characteristic for small values of r. These outcomes are influenced by the dimensionless time step based on the sensor depth, the spacing of the sensors along the surface relative to the sensor depth, and the amount of noise in the measurements, which was exact for this case. 6-3.2 Corrupted Data (Measurement Errors) The two-dimensional test cases are analyzed with measurement errors added to the simulated temperatures. Standard statistical assumptions are used to describe the measure- ment errors. The errors are modeled as additive, with zero mean, constant variance, and a normal distribution. A magnitude of approximately 1% of the maximum temperature rise 160 was selected as the standard deviation of the measurement errors. This error corresponds to 0.0018°C and 0.0025°C for the triangular and step heat flux, respectively. Estimation of the triangular heat flux with measurement errors is shown in Figure 6- 5a for a whole domain solution and Figure 6-5b shows a sequential solution with r=6. The two solutions are quite different, whereas the whole domain solution is relatively smooth, the sequential solution has a significantly larger variability. It is not clear if the variability of the SGM estimated heat flux is larger on time or on space. The sequential results are certainly unacceptable in comparison to the whole domain solution for r=6. a) GM Whole Domain 1 \ 0.8\ 0.6 w Heat flux (dimensionless) Figure 6-5a Estimated surface heat flux for triangular test case using data corrupted with measurement errors (0' = 0.0018°C ). Whole domain solution. 161 sing data corrupted with measurement errors (0' = 0.0018°C ). Sequential solution with r=6. Figure 6-5b Estimated surface heat flux for triangular test case u 162 Table 6-2 Estimation results for two-dimensional IHCP with simulated data corrupted with random errors Analysis domain Iterations A Tikhonov Comp S y as, per t1me Method r I (1.7- total seq int (sec) (0 C) W/m2 Triangular Flux (0' = 0.0018°C) - 11 Spatial Components GM all all 1.70 E-03 9 - 31.8 1.82 E—03 8.57 E-02 SGM 6 1 2.20 E-04 764 4.0 77.8 1.81 E-03 1.36 E-Ol SGM 8 l 4.50 E-04 758 4.0 96.0 1.80 E-03 1.02 E-OI SGM 10 1 7.00 E-04 746 4.0 1 19.0 1.80 E-03 9.19 E-02 SGM 15 1 1.25 E-03 718 4.0 173.7 1.80 E-03 8.65 E-02 SGM 20 1 1.50 E-03 759 4.3 246.9 1.80 E-03 7.53 E-02 SFSGM 6 1 5.00 E-05 767 4.0 77.1 1.81 E-03 1.57 E-Ol SFSGM 8 1 8.30 E-05 769 4.1 96.3 1.80 E-03 1.12 E-01 SFSGM 10 1 1.1 E-04 761 4.1 120.2 1.80 E-03 9.71 E-02 SFSGM 15 l 9.50 E-05 693 3.8 167.2 1.87 E-03 8.97 E-02 Step Flux (0' = 0.0025 ° C) - 11 Spatial Components GM all all 6.00 13-04 13 - 45.8 2.51 E-03 1.06 E—Ol SGM 6 1 1.90 13-04 775 4.1 73.3 2.50 15-03 2.18 13-01 SGM 10 1 5.5 E-04 762 4.1 121.9 2.50 E-03 1.24 E-01 SGM 15 1 7.5 E-04 822 4.5 200.7 2.50 E-03 1.15 E-Ol SGM 20 1 7.5 E-04 975 5.5 317.8 2.50 E-03 1.03 E-01 SFSGM 6 1 4.0 E-05 795 4.2 74.2 2.50 E-03 2.59 E-Ol SFSGM 10 1 2.0 113-05 953 5.1 150.8 2.50 E-03 1.49 E-01 Triangular Flux (0' = 0.0018°C) - 21 Spatial Components GM all all 1.70 E-03 9 - 54.7 1.84 13-03 8.50 E-OZ SGM 6 1 2.20 E-03 760 4.0 137.9 1.82 E-03 1.33 E-Ol SGM 15 1 1.25 E-03 718 4.0 328.2 1.82 E-03 8.44 E-02 SFSGM 6 1 5.50 E-05 786 4.1 141.6 1.82 E-03 1.65 E-01 SFSGM 15 1 9.50 13-05 693 3.8 314.4 1.88 E-03 9.29 E-02 163 Analysis of the triangular and step test cases with measurement errors are presented in Table 6-2. Numerical conditions similar to the exact cases were used, a dimensionless time step based on the sensor depth is At: = 0.06 with twenty-one nodes in the y-direction and eleven in the x-direction for the numerical solution. Data was assumed at two-hundred time steps and eleven components of the heat flux are estimated; twenty-two hundred heat flux components are estimated. The step heat flux case was more difficult to estimate. Reproducing a step change spa- tially is considerably harder than the more gradual changing triangular case. Figures 6-6a and b show the whole domain and SGM for r=15 estimated heat flux for the step case. Estimates have more variability than the triangular test case. Tabulated results in the mid- dle of Table 6-2 show that trends similar to the triangular case. The mean-squared error is as large as 25% of the maximum heat flux for small r-values. For the triangular case the mean-squared error did not exceed 16% of the maximum heat flux. The whole domain (GM) approach was superior to the sequential approach (SGM and SFSGM) for both two-dimensional test cases for r less than 10. This superior performance is supported by the smaller mean—squared error, final column of Table 6-2, and the shorter computational time, column seven. The mean-squared error was 7 to 83% greater for sequential approach, compared to the whole domain approach for the triangular heat flux, depending on the choice of r. Results were worse for the step heat flux, which had a mean- squared error that was 17 to 144% (depending on r) greater for the sequential approach. Computational times were adversely affected by the sequential approach. The sequential 164 a) GM Whole Domain f. ”if \ 7 / Z4 7 1.5\ 1\ 0.5\ 04 Ammo—coicoEEv 55 :33 Figure 6-6a Estimated surface heat flux for step test case using data corrupted with measurement errors (0' = 0.0025°C ). Whole domain solution 165 r=15 b) SGM 0.4 0.2 —— 22> 22K”— 22 22222 2 2222222225 2%%/Z§I £7.22: 7/j\\__ VVNK _ 2: ii: 2.2/ 2 2 (V/xmyfiifl 2% 2, 22222 i ..2 2 .225255 2222 .2222... .EE' 2/ :22 . Ti/ //4»Ml...! 2/ 15. 1 5 0. -052 15 93222022622223 23c 28E Figure 6-6b Estimated surface heat flux for step test case using data corrupted with measurement errors (0' = 0.0025°C ). Sequential solution r=15 166 approach required 60 to 400% more computational time, compared to a whole domain solution. The computational time increases proportional to the increase in the number of future time steps, r. The poor performance of the sequential approach for these cases can be attributed to two related factors. The first factor is that the sequential implementation is more ill-posed than the whole domain solution. Since the sequential solution solves over a shorter time domain, the effect of the measurement errors is more significant and results in a more ill- conditioned problem than the whole domain problem. Examining Table 6-2 shows that, even though the sequential problem is more ill-posed, the magnitude of the Tikhonov parameter in column four is decreased (except at large r—vales). Decreasing the Tikhonov parameter is contradictory to stabilize a more ill-conditioned problem. However to obtain the required magnitude in the sum-of-squares function, S y, the Tikhonov parameter must be decreased. The reason for this outcome (decreasing OLT for SGM) is the second factor in the poor performance of the sequential implementation. It is again the difficulty with the values near the end of the sequential interval influencing the early values. Much the same as was observed for the one-dimensional problem. Because the values near the end of the time region are biased, the regularization parameter must be decreased to provide the inverse solution “flexibility” to obtain the proper magnitude of the sum-of—squares func- tion. Although decreasing the Tikhonov parameter reduces the temperature sum-of- squares, it results in an increase in the variability of the estimated heat flux. In contrast to the one-dimensional solution, maintaining the heat flux constant over the sequential analysis interval does not improve the two-dimensional results. In the two- dimensional solution it does not work because the additional stability introduced by 167 maintaining the heat flux constant over the sequential interval requires a reduction in the Tikhonov parameter to obtain the desired sum-of-squares. The reduction in (X7- in turn increases the variability in the estimated heat flux. If r is too large when using function specification, the sum—of—squares can not be reduced to the desired level, regardless of the Tikhonov parameter. The estimated results are actually worse when maintaining the heat flux constant (with time) over the sequential interval for the two-dimensional problem. The sequential method performs poorly for small magnitudes of r, which are in the range that would typically be used for function specification. The estimated results for both test cases are improved as the number of future time steps is increased. Figure 6-7 demonstrates that the effects at the end of the interval are reduced by lengthening the sequential interval. At r=15 and 20 the mean-squared error in Table 6-2 is +1% and —12% of the mean-squared error of the whole domain estimation, respectively. The improved accuracy in the estimated heat flux using larger r-values is at the cost of a significantly increased computational requirement. Sequential solutions (SGM) for larger magnitudes of r require SOC-800% (depending on r) more computational time than the whole domain solution. Computational time increases proportional to the increase in r. However, the potential benefits of an on-line (real time) method may outweigh the increased computa- tional time. For the one-dimensional problem, and the two-dimensional problem with exact data, an improvement in the computational speed, with a comparable accuracy in the estimated function, was obtained when the heat flux was maintained constant over the sequential interval, the method was denoted SFSGM. In the two-dimensional analysis with measure- ment errors the results are not improved when the heat flux is maintained constant over the 169 sequential interval. Although computational times are comparable (to the SGM) for a specified r—value, the error in the estimated heat flux actually increases, as seen in the final column of Table 6-2, when maintaining the heat flux constant over the sequential interval. It was shown with exact data that a greater number of spatial components can be esti- mated than there are sensors available. The triangular case is reanalyzed for data with measurement errors and results are listed in the final block of Table 6-2. The outcome was similar to that seen in the analysis with exact data. A comparable heat flux, to that esti- mated with few spatial components, is obtained for a whole domain solution. Using the sequential method has a larger variability in the estimated heat flux for small r—values. The sequential method is improved by increasing r. Computational times are significantly greater using the sequential method. Several approaches are possible to improve the sequential method. Before outlining these approaches, a summary explaining the effects that are resulting in the sequential method’s poor performance is given. It is understood that the sequential method is more ill-posed. However, when solving with a sequential method the Tikhonov parameter is typ- ically smaller than the whole domain solution; as r is increased the Tikhonov parameter increases for the sequential method. The reason that the Tikhonov parameter decreases (assuming that the residual principle is applicable), even though the sequential problem is more ill-posed, is due to the components near the end of the time interval. These compo- nents are difficult to estimate; when using a gradient method with the adjoint equation approach, these components are not estimated. Instead, the adjoint function, which is pro- portional to the computed update to the heat flux, is defined to be zero at the end of the time interval. Consequently, the heat flux estimated at the end of the time interval is “near” 170 the initial estimate specified for the heat flux. Its deviation from the initial estimate depends on the magnitude of the Tikhonov parameter. As was demonstrated, lengthening the sequential interval improves the mean—squared error in the estimated heat flux; see Table 6-2 and Figures 6-7. Lengthening the sequential interval is also shown to increase the computational time in Table 6-2. An advantage of the SGM, compared to the SFSGM is that functional constraints are not imposed on the solu- tion. Consequently more than one component may be retained for a sequential interval and the computational time can be reduced by having fewer sequential intervals. For example, when retaining 10 components for the triangular heat flux with r=20, the computational time is reduced to 30.8 seconds, which is about only 3% greater than the whole domain computational time. The mean-squared error for this case, shown in the final column of Table 6-3, is within 1% of the whole domain error; the mean-squared has increased 14% compared to retaining only one component for r=20. Although lengthening the sequential interval and retaining more components on time improves the computational aspects, the process may reduce the likelihood that non-linear problems can be linearized. Since larger sequential intervals are considered, it is less likely that the changes in thermal properties, due to the temperature variation during the sequential interval, are negligible. This is an important advantage of the sequential method. The validity of this assumption will depend on magnitude of non-linearity in the problem. The one-dimensional problem was quite insensitive to the specified initial condition. Regardless of how the initial condition was specified, the solution required approximately the same number of iterations to converge. This same situation was found for the two- dimensional problem. Estimated heat flux and computational time were insensitive to the 171 procedure for setting of the initial condition for the two-dimensional problem. Whether the initial estimates for the next sequential interval were specified from the previous inter- val as outline in Section 5-4.2, or specified as zero, it did not significantly change the num- ber of iterations. A characteristic noticed for the one-dimensional problem, and two-dimensional test cases with measurement errors, was that the Tikhonov parameter decreases for the sequen- tial problem, compared to its magnitude for the whole domain solution. The decrease in the Tikhonov parameter was more pronounced at smaller r-values and approached the magnitude of the whole domain solution as r was increased. A smaller Tikhonov parame- ter was not expected for the sequential solution because the problem is more ill-posed. It is the sequential implementation that forces the Tikhonov parameter to be reduced, at least for small r—values. Since the sequential solution is on the shorter time interval, the zeroth order Tikhonov regularization is more influential because the sensitivity (or in this case adjoint function) is not as large, particularly for r-values beyond four to five time steps from the end of the interval, which represents a dimensionless time of 0.3. The nature of zeroth order regularization is to bias or “drive” the estimates to zero, to reduce their vari- ability. Consequently, to obtain the desired sum-of—squares, the Tikhonov parameter must be decreased so as not to “drive” the estimates towards zero. As r is increased to be farther away from this region the sensitivity (or adjoint function) is larger and the regularization has less affect on the solution. When solving in a sequential manner, however, there is additional information avail- able concerning the heat flux. The heat flux need not be “driven” to zero with the zeroth order regularization, instead it can be confined using the concept of prior information. The 172 prior information for the heat flux is specified from converged values at the previous sequential interval. The initial estimate for the sequential interval, which is specified from the converged estimates at the previous sequential interval, are set identically to be the prior information. The prior information is seen to enter the solution as q pn.(r, t) in equa- tion (4-2.4), which is the Tikhonov regularization term in the sum-of-squares function. Hence, instead of penalizing estimates that are different from zero, it penalizes estimates that are different from the previous estimates. For components that are far enough away from the end of the sequential interval the prior information has little influence. Near the end of the sequential interval, however, the prior information is more influential and helps control components in this region. The results using prior information are given in Table 6-3. As shown in the table, by using the initial estimate as prior information, the sequential method is improved com- pared to no prior information. Figure 6-8 shows the estimated triangular heat flux for r=6. Compared to the estimated flux without prior information, see Figure 6-5b, results are sig- nificantly smoother. For r=6 the mean-square error is reduced by 15%, and the computa- tional time is reduced in half with prior information. For larger magnitude of r the mean- squared error is comparable, while computational time is approximately one-half its former value without prior information. Similar trends are shown for the step heat flux case in Table 6-3. Notice that the magnitude of the Tikhonov parameter is greater when using prior information, but because it penalizes changes from the prior information (esti- mates at previous sequential interval) it does not bias the estimates as significantly. Prior information allows for smaller r—values to be considered and reduces computational time by requiring fewer iterations. 173 Heat flux (W/mz) b) SGM 1i / O.8\ , / /,/4///2 ‘ 0'6 ///’/////"//////% 0\ in"? ///i/////////// $1.3) :‘Ifl, '{itgig’g/{éEg/éflé Ia , I ’ ’i/ I I Figure 6-8 Estimated surface heat flux for triangular test case using data corrupted with measurement errors (6 = 0.0018°C ). Sequential solution with r=6 using prior information. 174 Table 6-3 Estimation results for two-dimensional IHCP with simulated data corrupted with random errors using prior information Analysis domain Iterations A Tikhonov Cpmp S y 0'5, per time o 2 Method r I “T total seq int (sec) ( C) W/m Triangular Flux (O’ = 0.0018°C) - 11 Spatial Components GM all all 1.70 E—03 9 - 31.8 1.82 E-03 8.57 E-02 SGM1 6 l 1.40 E-03 396 2.1 37.6 1.80 E-03 1.16 E-01 SGMl 8 1 3.50 13-03 391 2.0 50.1 1.80 13-03 1.02 13-01 SGMl 20 10 3.00 E-03 59 3.2 19.4 1.82 E-03 8.21 E—02 SGM 20 10 1.20 E-03 90 5 30.8 1.80 E-03 8.51 E-02 SFSGMl 6 l 8.30 E-04 390 2.0 36.5 1.81 E-03 1.33 E-Ol SFSGMI 8 1 2.50 E-03 384 2.0 48.9 1.82 E-03 1.11 E-01 Step Flux(O' = 0.0025 ° C) - 11 Spatial Components GM all all 6.00 E-04 13 — 45.8 2.51 E-03 1.06 E-01 SGMl 6 1 9.00 E04 508 2.7 48.2 2.50 E-03 1.79 E-Ol SGMl 10 l 3.50 E-03 423 2.3 67.8 2.51 E-03 1.30 E-01 SFSGMl 6 l 4.00 E-04 506 2.7 47.2 2.50 E-03 1.97 E-Ol SFSGM' 10 1 3.00 E-03 388 2.1 61.8 2.52 13-03 1.19 13-01 1 Prior information used for solution A final procedure to improve the sequential method is discussed. Although the proce- dure is not numerically studied, it has promise for further improving the sequential method. For all results presented herein, the Tikhonov parameter was specified to be con- stant for the entire time region. Using the residual principle, the correct magnitude of the 175 Tikhonov parameter was identified. This procedure allowed for a direct comparison of computational time for the whole domain and sequential methods. An alternative method could vary the Tikhonov parameter during the sequential analysis. In regions that the heat flux is rapidly varying, the Tikhonov parameter could be reduced to minimize the effect of bias. Similarly, it could be increased in regions where the heat flux is relatively constant. In a sequential implementation information from previous sequential intervals is available to modify the magnitude of the Tikhonov parameter, as well as the magnitude of r. The whole domain solution does not have the necessary information to select these parameters. 6-4.0 Results - Experimental Measurements The experimental configuration previously used to estimate the (orthotropic) thermal properties of the carbon-carbon composite is used to study the two-dimensional IHCP. Recall that transient temperatures and heat flux were measured to estimate the thermal properties in Chapter 3. Now the same transient temperature measurements, with the pre- viously estimated thermal properties, are employed to estimate the two-dimensional tran- sient surface heat flux. Surface heat flux is known for this experiment because the power to the heater is measured. Furthermore, the experiment is symmetrically designed and heat losses are negligible compared to the applied heat flux. Since the heater design has three independently controlled heaters, by activating only one of the three heaters a two-dimen- sional surface heat flux is experimentally prescribed. Additional details on the experimen- tal configuration are discussed in Section 2-2.0 and 3-2.0. An important issue in the study of the multi-dimensional IHCP is defining the spatial representation of the unknown heat flux. In the analysis of simulated data, it was shown that the gradient method could estimated more spatial components than the number of 176 sensors available. A rule used by the author is that the number of components estimated should not exceed the number of sensors. If more components than this are estimated, additional regularization is typically required to stabilize the solution. In the analysis of this experimental case, no assumptions on the spatial form of the heat flux are used. It is assumed that the heat flux is unknown only on the surface where the temperature sensors are located. All other boundary surfaces are insulated. See Figure 3-1. Hence, the heat flux that is estimated has a step change spatially and also a step change on time; a difficult case. The measured surface heat flux is shown if Figure 6-9a. Note that the surface with unknown heat flux, which is 7.62 cm long, has five sensors located above the heater at locations, x = 0.89, 1.91, 3.18, 4.45, and 6.73 cm. The numerical solution for this problem uses twenty-one nodes along the x-direction for all materials. A schematic of the geometry is given in Figure 3-2. In the y-direction there are four nodes across the mica heater, eleven nodes across the carbon-carbon, and eleven nodes across the insulation. A numerical time step of 0.64 seconds, which is the same as the measurement time step, is used. Because effective properties have been esti- mated for the mica heater, the interface between the heater and specimen is modeled with perfect contact. Although the mica heater is quite thin (0.44 mm), because its properties include the contact resistance between the heater and the carbon-carbon, it is thermally significant. Based on the effective properties of the mica heater, the dimensionless time step (Fourier number) for the temperature sensors is 0.23. A dimensionless time step in this range does not represent a difficult one-dimensional case; however, in two-dimensions with a limited number of sensors, the problem is more difficult. 177 6000 \ 5000 \ 4000 \ 3000 \1 2000 \ Heat flux (W/mz) 1000\ O\ 80 60 t( seconds) 40 Figure 6-9a Measured surface heat for experimental case to estimate thermal properties of carbon-carbon 178 Selecting the appropriate Tikhonov parameter requires knowledge about the errors to apply the residual principle (Alifanov, 1994). For this two-dimensional experiment the temperature was measured at seven locations, but two measurements are available at all seven locations. The sensors located at the back surface of the carbon-carbon composite provide little information compared to the sensors near the surface. Therefore the five sen- sors located closest to the heater are used in the analysis. By comparing the two measured signals at each location an indication of the noise in the measurements is gained. The stan- dard deviation of the redundant measurements, from the average, is computed as 1 J, 2 62(0) = {171—122 [Yk(t)—Yk,l(t)]2} (6-4.l) (=1 where Y_k(t) is the average temperature (for J k sensors) for sensor location k, Yk’ ,(t) is the measured temperature for sensor location k and sensor number I , and J k is the num— ber of sensors at location k. For this experiment J k = 2 for all k. Since so few sensors are available, by substituting for the average temperature Yk’1(t)+ Yk,2(t) Y_t = 6-4.2 k( ) 2 ( ) equation (6-4. 1) simplifies to l l 2 2 67*“) = {ifl’k’ 1“)" Yk,2(t)] } (6-4-3) Notice that this error is a function of time and sensor location k. An approximation of the total error is obtained by averaging the error in equation (6-4.3) over time and sensor loca— tion. The average magnitude is 0.1°C . This is about 20% larger than the measurement 179 error estimated for the one-dimensional experiment. The one-dimensional experiment had a larger sample set, J k = 7. The estimated heat flux is shown in Figure 6-9b and c for the whole domain and sequential method. The two methods are remarkable similar. Table 6-4 shows results from the analysis. The mean—squared errors are within 5% of the GM for all cases considered, except when prior information is used. For r=6, however, the computational time 77% greater for the SGM, which increases to 203% greater for r=10. The computational time is reduced to the level of whole domain solution by retaining more than one component. With r=6 and retaining two components, the computational time is cut in half and the mean-squared error is reduced by 3%, compared to the estimates while retaining only one component. Similar improvements are seen when retaining five components for r=10. For a comparison, another program for estimating the surface heat flux is used to ana- lyze the same data. The program QUENCHZD (Osman and Beck, 1989) computes surface heat flux from internal measurements using a combined function specification regulariza- tion method (CFSRM). In the CFSRM algorithm function specification on time and Tikhonov regularization on space is applied. The finite element direct solver TOPAZZD is used in this program. Heat flux estimated with QUENCHZD is shown in Figure 6-9d for r=6. The Tikhonov parameter was specified to have the same magnitude as the whole domain solution. Similar numerical aspects as used in the gradient method’s FCV solution were defined for the finite element solution. Approximately the same number of nodes and time step were specified for the numerical solution. The estimated heat flux is again quite similar. For the CFSRM only seven spatial com- ponents are estimated because of program limitations. Because QUENCHZD required two 180 b) GM Whole Domain / , . / 2 2 22222222222 2 2227:2222; 2722' 222,22 ’2.— 22,//22///2//// I 2222,, 2:222 , 2.22222. 22 .2221 22%/é l/! 2 297/why, 22' 722/2»; ,2- \1/ S d n O C e S ( t \\\\\\1W\ 00000000 mmmmmw m 6543211 1... @533 5e 28: Figure 6-9b Estimated surface heat flux for experimental case. Whole domain solution. 181 c) SGM r = 10 6000 \ 5000 \ 000 \ 2000 \ 2.2.53: 522. 28: t( seconds) 40 Figure 6-9c Estimated surface heat flux for experimental case. Sequential solution with 10. r: p—A 82 d) CFSRM 6000\ ' = 6 500m 4000. 3000\ 2000‘ 1000\ Heat flux (W/mz) t( seconds) Figure 6-9d Estimated surface heat flux for experimental case. Combined function specification regularization method with r=6. 183 Table 6-4 Estimation results for two-dimensional IHCP with experimentally measured heat flux Analysis domain Iterations A Tikhonov Cpmp S y Gq per time Method r 1 “T total seq in: (sec) (°C) W/m2 Experimental Case 1020&65 - 21 Components GM all all 1.5 E-06 l3 - 89.5 1.00 BO] 648 SGM 6 1 3.0 E-06 763 4.3 159 1.04 BO] 660 SGM 6 2 2.4 E-06 429 4.9 89.3 1.03 E-01 637 SGMl 6 1 1.3 E-OS 486 2.8 101 1.01 E-Ol 878 SGM 10 1 3.6 E-06 825 4.8 298 1.01 ED] 679 SGM 10 5 2.9 E-06 215 6.3 78.3 1.07 E—Ol 635 Experimental Case 1020&65 - 7 Components SGM 6 l 2.6 E-08 504 3.3 104 1.04 BO] 500 SGM] 6 1 2.0 13-07 371 2.2 76.8 1.00 E-Ol 389 CFSRM 6 l 1.5 E-06 - - 7200 7.88 E-02 399 CFSRM 10 l 1.5 E-06 - - 7200 1.65 ED] 518 1 Prior information used hours to obtain a solution the appropriate ocT, based on the residual principle, was not identified. It appears the appropriate (r, (1T) for CFSRM is between the two cases listed in Table 6-4. The estimated heat flux is not sensitive to these parameters. The biggest distinc- tion between the two programs is the computational time. The gradient method requires on the order of 100 seconds, while CFSRM requires on the order of 7000 seconds. These computational times can not be directly compared because different direct solvers are 184 used. The gradient methods employ a ADI method, which typically is much faster than the finite element method used in QUENCHZD. To understand the computational difference, the time required to obtain one direct solution with comparable numerical conditions was conducted. The ADI solution required 3.6 seconds while the FEM required 110 seconds; a factor of 30. Indicating that the gradient methods are more computational efficient because the inverse solution required a factor of 70 more computational time. Furthermore, the CFSRM only estimated seven components, compared to 21 for the GM. As more spatial components are estimated the CFSRM requires solving additional problems, some compu- tational savings are carried from previous solutions, however. In contrast, the gradient methods do not require solving additional problems when more components are consid- ered. A very good feature for the multi-dimensional problems. It appears that the gradient methods are more efficient for higher dimensional problems, but further cases, with simi- lar direct solvers, or more detailed studies are need. Comparing the gradient methods with standard function specification was not a thrust of this dissertation. Early indications are that the gradient methods are very promising for multi-dimensional problems. Chapter 7 SUMMARY AND CONCLUSIONS A study of multi-dimensional inverse thermal problems was presented. Two main problems were studied. Parameter estimation techniques were applied to characterize ther- mal properties of a carbon-carbon composite material. The second problem developed the sequential implementation of a gradient method, utilizing an adjoint equation approach, for estimating the surface heat flux from internal temperature measurements. Both prob- lems addressed multi-dimensional cases. Thermal properties, represented by two components of thermal conductivity and the volumetric heat capacity, were estimated for the carbon-carbon composite. One- and two- dimensional transient experiments were analyzed to estimate the thermal properties up to 500°C . Several conclusions are supported: 1. The thermal properties of the carbon-carbon are described by a parabolic temperature :C , and by a lin- dependence, given in equation (3-40) and (3—41) for the k; CC and (pC) ear temperature function, given in equation (3-37) for k: cc. 2. Experimental uncertainty in the estimated thermal properties was a maximum of 5.8% in (pC); and 2.4% in k1“. for one-dimensional experiments. Two-dimensional experimental uncertainty was a maximum of 6.0% in (pC); and k: CC and 4.4% in k8 X,CC' 185 186 3. One and two-dimensional results for (p C )2 c and k; CC correlate within the experimen- tal uncertainty. 4. An orthotropic model adequately describes the carbon-carbon material studied. A sequential gradient method was developed to solve the IHCP. Several one and two dimensional cases, using simulated and experimental measurements, were investigated to characterize the sequential method. A comparison between the sequential method and a whole domain solution was made. The following conclusions are drawn concerning the sequential gradient method, and variations of the method that were proposed: 1. A sequential gradient method has an accuracy comparable to the whole domain gradi- ent method, but in certain cases requires significantly more computational time. 2. In one-dimensional cases the minimum length of the sequential interval was a dimen- sionless time of 0.3. The dimensionless time is based on the depth of the sensor nearest the location of the estimated surface heat flux. 3. Applying function specification on time, in conjunction with the sequential gradient method, reduced the computational requirements and allowed for sequential intervals to be shorter than 0.3 for the one-dimensional cases. Computational requirements were comparable for the sequential function specification gradient method and whole domain method. 4. The two-dimensional cases required the dimensionless time to be greater that 1.0 to obtain comparable accuracy between the sequential and whole domain solutions. The dimensionless time is based on the depth of the sensor nearest the location of the esti- mated surface heat flux. 5. Applying function specification on time, in conjunction with the sequential gradient method, did not improve the computational requirements when errors were present in the data for the two-dimensional case. 6. Retaining more than one component for a sequential interval improved computational requirements without affecting the accuracy of the sequential gradient method for the two-dimensional case. 7. Using prior information with the sequential gradient method for the two-dimensional case was shown to improve the accuracy, permitted smaller sequential intervals, and made the computational requirements competitive with the whole domain solution. Chapter 8 RECOMMENDED FUTURE WORK In the course of this dissertation several topics for future work were realized. The top- ics are separated according to the parameter estimation or inverse heat conduction prob- lem. Experiments conducted to estimate the thermal properties of the carbon-carbon com- posite material measured transient temperature and heat flux histories for discrete experi- ments spanning the desired temperature range. Experiments with one- and two- dimensional heat flow were analyzed independently, assuming the thermal properties were constant for the analysis of each experiment. In the end, the discrete experiments were combined to determine a temperature dependence for the thermal properties. One and two- dimensional results were also compared. Topics of future direction are noted: 1. Analysis of currently available data to directly estimate temperature-dependent thermal properties is of interest. A sequential analysis using prior information is used to accom- plish this. 2. The optimal experiment for estimating two components of thermal conductivity and volumetric heat capacity is needed for the case of a relatively high thermal conductivity material (which this particular carbon-carbon has). The optimal conditions for the design of a series of experiments is an important concept that needs attention. The experimental series may included one- and two-dimensional experiments. As the com- plexity of the experiments increase, more frequently several experiments may be required to fully characterize complex materials or compound structures. 3. The model of the carbon-carbon composite can be extended to include the silicon-car- bide layer. This case of a thin layer on a thick substrate represents an important prob- lem. Work to estimate the thermal properties of the thin layer, in conjunction with the underlying material, is of interest. 187 188 . Methods, both experimental and analytical, should be generalized to the three-dimen- sional case. Developing the analysis is straight forward. Experimental techniques, including the experimental design, are extremely important for higher dimensions and will require significant effort. The sequential gradient method has several areas for potential growth of the method. . The procedure to linearized nonlinear problems in the sequential gradient solution sug- gests large savings in computational time without significantly affecting accuracy, the procedure needs investigating. . Implementing on-line logic for selecting the regularization parameter, number of future time steps, and/or the number of retained components would enhance the sequential method. Quite frequently, the selection of these parameters is not trivial and requires experience and an understanding of inverse problems. Since the accuracy of the solu- tion may be sensitive to correctly selecting these parameters, the method would be enhanced to logically select the parameter. . In the sequential formulation presented, zeroth order Tikhonov regularization was used. Extending the method to include first order Tikhonov regularization is appropriate. . Prior information to spatially filter the heat flux could provide spatial regularization. 5. Implementing the methods for the three-dimensional case is of interest. APPENDIX APPENDIX A FINITE CONTROL VOLUME METHOD A-1.0 Introduction Many methods exist for discretizing partial differential equations. Two popular meth- ods are finite differences and finite element. An alternate approach applies conservation of energy directly to a control volume. That method is applied to discretize the heat conduc- tion equation here. The problem considered is the heat conduction equation with constant thermal prop- erties, a spatially time dependent volume energy source 8 er a 87' _ er a 5:0“ 5;)+$(kt5;)+g(x, y, t) — pC[§; +u0§£| (A-1.l) 8T _ ._ -k é—n—i — qi(si, t) (r --l, 2, 3, 4) (A-1.2a) F, OR Tlr = Ti(51" t) (i=1, 2, 3,4) (A-1.2b) T|t=0 = T0(x, y) (A-1.2c) The thermal conductivity is assumed orthotropic. First or second kind boundary condi— tions are allowed on the four bounding surfaces (1“,) . Conditions may have temporal and spatial variation. The outward normal of surface 1' is ni. Since a regular geometry is used this is the unit normal in the x or y direction. 189 190 The FCV procedure uses energy conservation principles to derive a discretized set of equations to solve for the temperatures. In contrast, a finite difference scheme would dis- cretize the problem presented in equation (A-l.1) and (A-l.2), i.e. use a finite difference approximation for the derivatives, to obtain a numerical solution. The FCV procedure is preferred over a finite difference approach because it is more general and applies energy conservation principles to derive the equations. To implement the FCV procedure the spatial domain is discretized with uniform mesh points (nodes) (x, y) = (iAx, jAy) i: (o, 1, 2 ..... ,1), j = (o, 1, 2 ..... ,J) (11-1.3) where the mesh lengths are L L, ( = .5 (Ay = 72) (A-1.4) Control volumes are the region (AxAy) surrounding the nodes with the node centrally located. The control surfaces are at the mid-plane between the nodes. Figure A-l shows a schematic of the nodes, control volumes, and control surfaces. The integral energy equation (Beck et al., 1992) J(_q . fi)dA + Jgdv = Jnggdv+ Iqu’ - fi)dA (A-1.5) C.V C.S. C.S. C.V. is applied to each control volume on the domain to obtain the FCV equations. The integral energy equation can be used to derive the describing differential equation in equation (A- H) as well. The energy balance for a typical node in the interior is shown in Figure A-2. Using the simplified approximation of equation (A-1.5), which assumes the quantities are 191 Control _>| Ax I‘— Surface “Ode l _I___L__L__|____l__l___l __I____I_.__I_ | I I I | I | l I I I O I O I O I O I O I O I O I O I O I -—|———I-————I——-—I—-—-I---I--—-|———|—-—-I-——I— Ay i1.1.1 I I('5J)I 11.1.1 I | l I I | | l | I ‘l"—‘T-—1"”|_—'T“"T'__|-—"I—“T—"'I— T 1 I o I o I o I o I o I o I o I o I o I I I I I I I | I l I '"I—__I-_T—_I__—I__T__|_—_I—_T_-I- j=O I. 1 i=0 1 2 ..... I Figure A-l Computational nodes and control surfaces (1.131) ‘11. u I'__f--‘I u (i_1,j)o I o —I-—> o(i+l,j) ‘11: I (131’) I 41-. .__I___. ‘11- o (inf—'1) Figure A-2 Typical energy balance for an interior node 192 constant over the control volume and surface, gives 3T _ (_ 417411 + qi+Ai+ _ 411"}: + qj+Aj,) + gV = pC-é; V + pvoui+Ai, — pvoui-Ai. (A-1.6) The heat fluxes are taken as the mean values over the control surface and are approximated as (Ti,j_ Ti- l,j) ‘11; = _kx Ax (A-1.7) qr, = -kx(Ti+1X; TU) (A-1.8) ‘11. = ‘kym’flaly- TU) (A-L9) Assuming that the velocity is positive, the flow terms are evaluated at the upwind locations to avoid the known difficulties associated with central differences for this term (Anderson etal.,) u- = C T«,- (A-l.11) “i- = C1971- (A-1.12) l,j The areas are (A1.) = Ai = Ay) and (Aj. = A]. = Ax) and the volume is (AxAy). Substi- tuting equation (A-l.7) through (A-l.12) with areas and volume into the energy balance gives (Tij—Ti,j—l) (Ti.j+I_Ti.j) Ay Ax+ky Ay (Ti.j—Ti~l.j) (Ti+1,j"Ti,j) k1 Ax A_v+kx Ax Ay — kv Ax 8T. . +gAxAy = pCAxAy-a—;"1+pCv0Ay(Tl-vj— T,-_ UNA-1.13) 193 Defining the following constants k,r k x, = —; x = ——-”—— (A-l.14) PCUM)2 y PC(A)’)2 and simplifying equation (A-1.13) results in an equation that represents an energy balance on a finite control volume V V [(26, + A—QT" 1,1- (226x + icy“ + it'xTi+ 1, j] + [4.3713 j_ 1 — 2231131.... MT“). 1] + iLé = g—tTi’jm-LM) For boundary conditions that are prescribed temperature only, after approximating the time derivative in equation (A-1.15) the equation can be written for each interior node and numerically solved for the all discrete temperatures. However, if flux boundary conditions are prescribed, a FCV procedure performs an energy balance on the boundaries and results in an equations to describe the energy balance there. The FCV equations to represent the boundary conditions apply an energy balance at the boundary FCV. Two possible configurations on the boundary are shown in Figure A-3. The first is along a bounding surface but not on a comer and the second is at a comer. The energy balance for the bounding surface node is 2 3T —(" qi+Ai+—qjjAj_+Qj+Aj++qsth5)+gV = 9C3? V+pv0u0Ai—pvou1A,-+ (A’l°l6) The heat fluxes (q 1' , q j, q j ) and flow term (ui ) are the same as for an interior grid point and given in equation (A-l.8) through (A-1.10) and equation (A-l.l 1). The flow term at the boundary surface is evaluated using the average temperature, ui = Cp(T0,j+Tl,j)/2' The areas are A}. = A}: = Ax/2, AI. = A5 = Ay and the 194 0, '+1 0 ( J ) q” (0, 1) J o I' _ 1 q" ui<—I Ii? "Lb:- — 11" u!” . 4—5 —I;I> 00’!) 3132,,- I ' €1,- qsw‘ (0,J')| | 1* (0,0)‘ “ — " * ' ~ I I . (1,1) L _ J qsl q}.- (09 j_ 1) . a) Bounding surface node b) Comer node x=0 (x=Qy=0) Figure A-3 Finite control volumes along the boundary of the domain volume is V = AxAy/Z. Substituting these relations into equation (A-l .16) and simplify- ing gives 1 V0 , V0 1 1 ' [PMEITOJ+IZMEITLJI+ [1,181-1 -2x,r,,+>.,r,j, 11 2q52vj g0,j _ a_TOvj —pCAx pC — a: (A-1.17) where Xx and Xx are given in equation (A-12). The prescribed heat flux (“15" j is the heat flux on surface 32 at nodej. Using the same procedure the following equation is derived to represent the energy balance at a corner node 195 V V [(423, + $970, 0 + (21', _ E1)“ 0] + {—271370, 0 + 223,10, 11 2631.1“ ‘I’ 2‘49: IA)’ 81,; 3T1.) — + —- = — (A-1.18) pC(AyAx) pC a: The equations derived for the boundary conditions, equation (A-l7) and (A-18), apply along boundary x = 0 and at comer (x = 0, y = 0). Similar expressions can be derived for the remaining boundary surfaces and corners. A-2.0 Alternating Direction Implicit Method (ADI) Discrete equations derived can be solved using an explicit or implicit method. The explicit is more easily implemented, however, may be unstable depending on thermal properties and the time and space discretization. In contrast, an implicit method is more complicated to implement, but unconditionally stable. A fully implicit method has five unknowns for each interior node (see Figure A-2 and equation (A-1.15)) and requires solving a penta-diagonal system for each time step. A more efficient method is the ADI (alternating direction implicit) scheme, which requires solving two tri-diagonal system of equations at each time step. The time domain is uniformly discretized t=nAt n=(l,2 ..... ,N) (A-2.l) In the ADI method, two solutions are obtained for each time step. For the first half time step the FCV equations are solved implicit in the x-direction and explicit in the y- direction. For the second half time step the difference equations are implicit in the y- direction and explicit in the x-direction. Hence, the ADI method is a two step solution. An 196 intermediate solution at time (n +1/2)At is obtained after the first step. The second results in the solution at the end of the time step. A-2.l ADI Equations (n+1/2) The FCV equations using an ADI scheme proceed over the first half time step. These equations are implicit in the x-direction and explicit in the y-direction. For illustrative pur- poses consider the FCV equations in equation (A-l.13), (A-1.15), and (A-l.16). Introduc- ing the forward difference for the time derivative and the implicit/explicit ADI scheme for the first half time step produces the following equations Interior Node: (0