t.>lo'.>ltx K run-{Way}. 2c .7? L E u... am?“ Mr , s .. u. «Fem ‘ 34,33... : \ A».L.%l .h «mm .M ‘ C. . . 3 3...... ‘ it... Q V 5-7 J I‘vw L | 0‘ ‘ ‘5. ffimm . was. run. at: figvflfi I3 ; i ..(r. x. 15.3".” 1006 This is to certify that the dissertation entitled PARAMETER ESTIMATION IN MULTI-AXIAL THERMAL DIFFUSIVITY EXPERIMENTS presented by Sean Edgar Davis has been accepted towards fulfillment of the requirements for the Doctoral degree in Mechanical Engineering ”$44!”. Origlll/ Majod‘ o ssor’s Signature 1 J03. 1’0; Date MSU is an Affirmative Action/Equal Opportunity Institution - _....-o-.-.-.-.-.--—.-.-._-._._ _._ V._ . LIBRARY Michigan State University PLACE IN RETURN Box to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 2/05 c:/CIRC/DateDue.indd-p.15 PARAMETER ESTIMATION IN MULTI-AXIAL THERMAL DIFFUSIVITY EXPERIMENTS By Sean Edgar Davis A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Mechanical Engineering 2005 ABSTRACT PARAMETER ESTIMATION IN MULTI-AXIAL THERMAL DIFFUSIVITY EXPERIMENTS By Sean Edgar Davis Thermomechanical analysis requires, in general, quantifying the thermal response of materials and thus, the therrnophysical properties of thermal conductivity (or diffusivity) and specific heat. Many materials yield an anisotropic thermal response, requiring measurements of multiple components of the thermal conductivity tensor. The extended flash method allows simultaneous measurement of multiple components of the thermal diffusivity tensor along any axes of interest. The locations of the temperature sensors in such an experiment have an affect on the ability to accurately estimate the desired components of the diffusivity tensor. Prior studies have optimized the sensor geometry for specific cases, but a general methodology for optimizing such an experiment while allowing flexible positioning of the sensors during measurement has not been developed. Here, D-optimization is applied to a simulated extended flash diffusivity experiment to improve the accuracy of the extended flash method through optimization of the distance between the sensors. The effects of experimental noise and heating power and duration are also examined. Optimal sensor positioning was determined using two criteria: maximizing the information obtained via all the estimated parameters, and maximizing the amount of information obtained via the estimated diffusivities. Results indicate that the optimal inter-sensor distance increases with an increasing ratio of in-plane to out-of- plane diffusivity. The analytically determined optimal sensor positioning for an isotropic material is validated via experimental measurements on A181 304 stainless steel, where it is shown that the accuracy of the estimated parameters improves for data sampled at the optimized locations. When modeling the anisotropic thermal response of materials, the material may be rotated such that the physical axes coincide with the principal axes of the thermal diffusivity tensor, resulting in thermal orthotropy. During measurements of such a tensor, however, the principal axes may be unknown, requiring a method to determine principal values and the orientation of the principal directions while simultaneously measuring the diffusivity. An analytical study was performed where the four non-zero components of the diffusivity tensor a were estimated for a material possessing random in-plane anisotropy on the order of certain manufactured or mechanically loaded elastomers. Results indicate that a four-sensor array allows sufficient sampling of the material response to permit estimation of a to within 1% of the reference values. When orthotropy is assumed for a material exhibiting random in-plane anisotropy, the estimated values of on, are resolved to within 0.4% of the reference values when the magnitude of the anisotropy is on the order of that seen in common deformed or machined elastomers. Should the orientation of the specimen be known prior to experimentation, it is shown that sampling the temperature response in an orientation that is rotated from the principal thermal axes may provide more information about the estimated parameters, thus improving the accuracy of the experiment. ACKNOWLEDGMENTS Foremost, I would like to thank the members of my Ph.D. guidance committee, Drs. Neil Wright, James Beck, Tom Pence, and Andre Lee, for their input and assistance during the preparation of this dissertation. I would particularly like to thank Dr. Wright, my advisor, for his encouragement and patience. I hope that his attention to detail, something we both admire, is reflected in my preparation of this work. I am grateful for the support of the National Science Foundation and the Whitaker Foundation, both of which generously provided funding that made this research possible. I would like to thank Xin Huang for her help in conducting experiments at Michigan State. I owe Dr. Edwin LeGall, a former colleague in Dr. Wright’s lab, thanks for his encouragement in pursuing a Ph.D. I have made many friends during my time here at Michigan State; these friends have provided relief when difficulties were encountered in this research and their aid is gratefully acknowledged. Finally, to my dear friends back in Baltimore who always made me feel like I was home despite moving hundreds of miles away, I thank you for your support. iv TABLE OF CONTENTS LIST OF TABLES .............................................................................. vi LIST OF FIGURES ............................................................................ vii NOMENCLATURE ............................................................................ x CHAPTER 1. INTRODUCTION ............................................................... 1 CHAPTER 2. OPTIMAL POSITIONING OF SENSORS FOR THERMAL DIFFUSIVITY MEASUREMENTS .......................................................... 7 2.0 Motivation .............................................................................. 7 2.1 Methods .............................................................................. 15 2.2 Results ............................................................................... 24 2.2.1 Measurement Duration ................................................ 24 2.2.2 Anisotropic Diffusivity ............................................... 28 2.2.3 Experimental Validation ............................................. 31 2.3 Discussion ........................................................................... 37 2.4 Closure ............................................................................... 45 CHAPTER 3. ON MEASURING THE THERMAL DIFFUSIVITY TENSOR WHEN PRINCIPAL AXES ARE UNKNOWN A PRIOR! .............................. 46 3.0 Motivation ........................................................................... 46 3.1 Methods .............................................................................. 56 3.2 Results ................................................................................ 63 3.3 Discussion ........................................................................... 73 3.4 Closure ............................................................................... 79 APPENDICES .................................................................................... 80 Appendix A - Notes concerning IR thermography .................................. 81 Appendix B - Alternating Direction Implicit (ADI) program for modeling orthotropic heat conduction in a slab specimen ............................. 88 Appendix C — Random in-plane anisotropy finite difference program ........... 128 REFERENCES .................................................................................. 152 v LIST OF TABLES Table 2.1 Estimated directional thermal diffusivity values and associated standard deviations for a 1.50 mm stainless steel plate (average of four tests). 36 Table 2.2 Average deviation of the out-of—plane and in-plane diffusivities from the reference value of CL = 3.95 x 10'6 m2/s for a 1.50 mm stainless steel plate. 36 Table 2.3 Values of A and A+ computed from measurements of the 1.50 mm steel plate. 36 Table 2.4 Optimum sensor location based on A and A“ for various 7 in a phantom elastomer with an out-of-plane diffusivity of 1.08x10'7 mz/s and tp = 1.0 s. 39 Table 2.5 Relation between t,,, optimum A and N, and variation in 0t. 40 Table 2.6 Relationship between noise level in measurements and the corresponding accuracy in the thermal properties of interest for an isotropic specimen with tp = 1.0 s. 42 Table 3.1 Representative values of B, on 1, (122, and an used in the analytical study. 61 Table 3.2 Error in estimating diffusivity when assuming orthotropy (i.e., neglecting 0112) for an anisotropic material with principal axes oriented by B with respect to the measurement axes. The maximum shear diffusivity for this material is about 18% of the median axial diffusivity value. 69 Table 3.3 Error in estimating diffusivity when assuming orthotropy for an anisotropic material with principal axes oriented by B with respect to the measurement axes. The maximum shear diffusivity for this material is about 35% of the median axial diffusivity value for tests 18-22 and about 64% of the median axial diffusivity value for tests 23-27. 69 vi LIST OF FIGURES Figure 2.1 Diagram of the top surface of the slab specimen illustrating the illumination area and the in-plane geometry. The quarter plane modeled in the parameter estimation scheme is bounded by the positive x- and y-axes, which themselves are axes of symmetry. Typically, specimens are between 1.5 mm and 3.5 mm thick. Three sensors are located on the bottom surface of the specimen along the coordinate axes a distance d apart; the projections of these sensor positions are shown on the modeled quadrant of the specimen. Figure 2.2 Simulated temperature histories and inverse model reductions for an isotropic specimen at sensor locations of d = 0, 7, and 20 mm: (a) tp = l s, (b) tp = 10 s. O is defined as T-Ta, where T... is the ambient temperature. Figure 2.3 Representative residuals, defined as the difference between the modeled and measured temperature histories, for an isotropic specimen: (a) center temperature sensor, (b) off-axis temperature sensor. Figure 2.4 Panel (a) shows the in-plane temperature response and 11,- for an isotropic specimen where d = 12 mm. The dots represent the in-plane temperature data, the solid line the ax sensitivity coefficient, the dotted line the heat flux sensitivity coefficient, the short dashed line the 0tz sensitivity coefficient, the long dashed line the 0ty sensitivity coefficient, and the dot-dash line the convection sensitivity coefficient. Panel (b) shows A (circles) and A+ (triangles) as a function of elapsed time for the same specimen and sensor location. Figure 2.5 In-plane temperature response (circles) and log A” (triangles) as functions of time for an isotropic specimen with d = 12 mm. Figure 2.6 Variation of A and A“ with respect to d for an isotropic specimen with t,, = 1.0 5. Calculated values of A are represented by circles, while calculated values of A” are represented by triangles. Figure 2.7 Location of the maximum A (circles) and A+ (triangles) as a function of 7. Figure 2.8 Variation of A and A+ with respect to d for an isotropic steel specimen with t,, = 1.0 5. Calculated values of A are represented by circles, while calculated values of A+ are represented by triangles. vii 16 25 26 29 3O 32 33 35 Figure 3.1 Isotherms on the bottom surface of a specimen at t = 30 5 following 1 s heating. (a) transversely isotropic: an = an = 0.16 mmzls, on; = 0, (b) anisotropic: om = an = 0.16 mmz/s, an = 0.0282 mmzls, (c) orthotropic: a1. = 0.188 mmzls, an = 0.132 mmZ/s, on; = 0. Note that Figure 3.1c is congruent to Figure 3.1b if rotated 45° counterclockwise. 49 Figure 3.2 Schema of the top surface of the specimen illustrating the illumination area. Top and bottom faces are subject to specified constant and uniform convection coefficients. Heating on the top face is of known power and duration; lateral edges are adiabatic. The temperature sensors are located on the bottom surface of the specimen at the comers of a 9 mm square that is aligned with the coordinate axes and has one sensor at the center of the bottom face; the projection of these sensor positions are shown on the top surface of the specimen. The principal axes of a, x’ and y’, are oriented at an angle B from the experimental axes; B is assumed unknown at the start of parameter estimation. 58 Figure 3.3 Simulated temperature excess, (-3 = T-Tm, and inverse solutions for (a) orthotropic test (case 1) and (b) maximum shear test (case 5). Note that for case 5, the x- and y-sensor temperature histories are identical as a result of an and (1.22 being equal. 64 Figure 3.4 Scaled sensitivity coefficients for case 1: (a) y- sensor, (b) comer sensor, (0) center sensor, ((1) x-sensor. The solid line represents F for (133, the dotted line F for on 1, the dashed line F for an, and the dot-dash line F for an. Time is not scaled, as the choice of appropriate scaling parameters is not immediately obvious. For example, if the specimen thickness and corresponding axial diffusivity cm are used as scaling parameters, than 1 second would correspond to a Fourier number of 0.0104; however, if either of the principal in-plane diffusivities are used with the specimen width as scaling parameters, 1 second would correspond to a Fourier number of either 8.23 x 10'5 or 1.18 x 104. 65 Figure 3.5 Scaled sensitivity coefficients for case 5: (a) y- sensor, (b) comer sensor, (0) center sensor, ((1) x-sensor. The solid line represents F for cm, the dotted line F for on 1, the dashed line F for an, and the dot-dash line F for (x12. 66 Figure 3.6 Input data and model fits at the x-, y-, and comer sensors for case 5 (B = 45°) where orthotropic behavior is assumed during parameter estimation. The solid lines represent the model fits for the simulated temperature history at the indicated sensor. The poor fit between the data and the model temperature histories at the corner sensor is due to the assumption of no shearing diffusivity. The temperature responses at the x- and y-sensor locations overlap since the x- and y-diffusivities are equal. 71 viii Figure 3.7 Normalized A as a function of measurement duration for cases 1 through 5. Empty circles represent data taken for B = 0°, triangles B = 15°, squares B = 22.5°, diamonds B = 30°, and filled circles B = 45°. 72 Figure 3.8 Variation of A for an (I having principal values of [1.10, 1.88, 1.32] x 10'7 mzls. The amount of information obtained from a measurement increases as the specimen orientation shifts away from the principal axes of a. Note that for B = 0° and 25°, A is six and two orders of magnitude smaller, respectively, than A at B = 45°. 78 ix FOi hb h: Pi qo NOMENCLATURE specific heat capacity (J/kg K) intersensor distance (mm) Fourier number in the i-direction (= (It/[2) (ND) bottom surface convection coefficient (W/m2 K) top surface convection coefficient (W/m2 K) identity matrix (ND) thermal conductivity tensor (W/m K) component of the thermal conductivity tensor in the ij-direction (W/m K) halfwidth of heated area (m) halfwidth of specimen (m) number of sensors number of data points estimated parameter (varies) heat flux (W/mz) temperature (°C) glass transition temperature (°C) temperature measured at time j (°C) ambient temperature (°C) the maximum rear surface temperature excursion (°C) time (s) heating duration (3) experiment duration (5) thermal diffusivity tensor (m2/s) thermal diffusivity component in the i-direction (mZ/s) angle between the physical (experimental) and principal thermal axes of 0t (degrees) scaled sensitivity coefficient at sensor position i with respect to estimated parameter j (K) anisotropy ratio (ND) specimen thickness (m) Fisher Information Matrix (ND) density (kg/m3) standard deviation of thermal diffusivity (mz/s) temperature excursion, T-T... (K) chi-squared error (K2) determinant of the Fisher Information Matrix (ND) ratio of the determinant of the Fisher Information Matrix for all estimated parameters to the determinant of the Fisher Information Matrix for the nuisance parameters (ND) del operator (m") xi CHAPTER 1 INTRODUCTION As computational technology advances in fields such as engineering and medicine, quantifying material thermophysical properties becomes increasingly important. The response of many materials may be assumed to be isotropic; however, others such as tissues, composites and polymers may exhibit anisotropic responses. This anisotropy could be due to the material structure (e.g., crystalline or composite), manufacturing processes (Doss and Wright, 2000) or deformation (LeGall, 2001). Doss and Wright (2000), among others, have demonstrated that the thermal diffusivity in seemingly isotropic polymers may vary by 30% or more. While temperature is a scalar, thermal conductivity is, in general, a tensor. For such a material, the three-dimensional heat diffusion equation may be written as pc%f—=V(kVT) (1.1) where p is the density of the material in question, c the specific heat, T temperature, I time, and k the second-order thermal conductivity tensor. The thermal conductivity tensor has nine components k1 1 kn k13 k: k21 k22 k23 (1.2) k31 [<32 1‘33 however, since k is a symmetric tensor (which follows from classical thermodynamics and Onsager’s principle of microscopic reversibility (Onsager, 1931)), only six of the components are independent (e.g., k23 = k32) (Powers, 2004). Commonly, k is assumed to be isotropic, with k = k-I. In general, k may be a function of T; however, here only small temperature changes are considered such that k is a function of direction only. For many circumstances, anisotropy is considered such that the principal axes of k are coincident with the physical axes of the problem; then, only k; 1, kn and k33 have nonzero values. The principal axes of the thermal response are unknown for materials of unknown or complex composition, requiring the off-axis terms to be considered. The thermal diffusivity a, where a = (l/pc)-k, is often used instead of k in transient problems, and is often easier to measure. Possible reasons for anisotropy in polymers include reorientation of the molecular chains or the scission and reforming of inter-chain bonds during loading of elastomeric materials (Wineman and Min, 2003). Manufacturing processes of many plastics and elastomers create anisotropy due to molecular reorientation during molding, calendaring, pressing, and similar processes. Kilian and Pietralla (1978) illustrated thermal anisotropy in uniaxially drawn polyethylenes, while Broennan et al. (1999) measured anisotropic conduction in stretched silicone rubber. LeGall (2001) has shown that the thermal diffusivity of certain elastomers change during finite biaxial deformation. The orientation of the principal axes of thermal diffusivity may be changing during the deformation of a material, especially during general deformations, thus it cannot be assumed, in general, that the principal axes of conduction coincide with the principal axes of deformation. The challenge is to develop a procedure that provides optimal assessment of the magnitude of thermophysical properties of materials, as well as the direction of the principal axes of these properties. Methods to measure thermophysical properties must be robust enough to deliver accurate and precise results while providing indication of the accuracy of the model on which they are based. The estimation of thermophysical properties for a material involves two processes: a physical experiment in which a material specimen is heated and temperature history data is recorded, and a parameter estimation routine where the data are matched to an analytical model of the physical experiment, allowing the determination of the specimen material properties, often through an optimization scheme. Traditional experimental methods of thermophysical property measurement can be divided into those for thermal conductivity, usually steady state (e.g., the guarded hot plate; Bradley, 1943), and those for thermal diffusivity, usually transient (e.g., the flash method; Parker et al., 1961). Many methods require excellent contact with the material specimen and, thus, are unsuitable for tests during which the specimens may be deformed. Contact with physical temperature sensors such as thermocouples disturbs the temperature profile of the specimen, further complicating the problem (Keltner and Beck, 1983). The self-heated thermistor, originally proposed by Chato (1968), has been primarily used to measure thermal conductivity. Valvano and others have extended its use for thermal diffusivity measurements as well as perfusion in biological materials (Valvano et al., 1985; Valvano and Nho, 1991; Yuan et al., 1993), but the derived properties are volume averaged and thus do not allow measurement of thermal properties along axes of interest. Conversely, non-contact thermo-optical methods are often used to measure directional components of the thermal diffusivity tensor. These methods include the flash, thermal wave, and transient gradient methods. While the original flash method, developed by Parker et al. (1961) and standardized by ASTM (E 1461-92), only measures the component of thermal diffusivity perpendicular to the plane of heating, extensions of the flash method allow simultaneous measurements of diffusivity along any axes of interest (e.g., Maillet er al., 1990; Graham et al., 1999) and may even permit the measurement of off-axis diffusivity terms as well. A fully populated a tensor may be thus resolved with a single test, assuming a properly designed experiment. The parameters that are calculated in a parameter estimation routine may be broken down into two categories: the parameters of interest, such as thermal diffusivity, conductivity, or specific heat, and nuisance parameters, which are parameters that are not of interest to the experimenter yet must be estimated in order to arrive at accurate estimates for the parameters of interest (Beck and Arnold, 1977). Typically, nuisance parameters involve boundary conditions and may include an imposed heat flux, convection coefficients, or emissivities of high-temperature bodies. Ideally, the experimental design should render solutions for the parameters of interest relatively insensitive to errors in either the nuisance parameters or the geometric parameters of the experiment, such as sensor location. Only reasonably accurate estimates of the nuisance parameters may be needed in order to determine the parameters of interest to a high degree of confidence, thus, less computational effort may be expended in finding the nuisance parameters (Beck and Arnold, 1977). Location of the temperature sensors, however, often plays a significant role in the accuracy of the experiment (for an illustration of the relative effect of sensor positioning on the accuracy of thermal property estimation, see Mzali et al., 2004). The selection of temperature sensor location, specimen size, and heating geometry and duration a priori requires knowledge of the thermophysical properties to be measured. Of these variables, however, only temperature sensor location may be optimized along with other parameters post-experiment by measuring the temperature field history. The use of geometrically fixed sensors prohibits redesigning the physical experiment “on the fly”, so each time the optimization criteria indicate that the sensor locations should be moved, the experiment must be performed anew with the appropriate adjustments made to the sensor locations. Optical methods of temperature measurement such as focal plane array IR cameras can overcome this limitation by allowing flexibility in the choice of sensor position. The temperature sensitivity of available cameras is on the order of that of typical commercial thermocouples, and sampling rates now typically exceed 60 Hz for 320 x 256-pixel detectors. Most importantly, IR cameras allow full-field measurements on the specimen surface, although such field measurements yield large data files. Adding this level of optimization to the thermophysical property estimation routine increases the complexity of parameter estimation, but careful design of the experimental boundary conditions and the data reduction model can yield methods whose reduction time is comparable to previous non-optimized models for thermal property estimation. The use of such full-field measurements allows the experimenter to redesign the experiment to make best use of the information available to capture unknown components or orientations of the thermal diffusivity tensor 0. Presented here is (a) an optimization of sensor position and heating magnitude and duration in an analytical extended flash diffusivity experiment performed upon a material exhibiting transverse isotropy, (b) experimental verification of the analytical procedure using a steel plate, (c) an investigation of the ability to measure the four nonzero components of the a tensor in a specimen exhibiting monoclinic (in-plane) anisotropy where the orientation of the principal axes of a is assumed unknown prior to testing, and (d) quantification of the error inherent in the assumption of orthotropic behavior for a specimen that actually behaves in a monoclinically anisotropic fashion, i.e., the orientation of the principal thermal axes in one plane (where the behavior may be characterized as orthotropic) is not aligned with the physical axes of the experiment. CHAPTER 2 OPTIMAL POSITIONING OF SENSORS FOR THERMAL DIFFUSIVITY MEASUREMENTS 2.0 Motivation Advances in thermomechnical analysis of polymers, composite materials, and biological tissues are leading to improved designs in manufacturing and medicine. Thermomechanical analysis requires, in general, quantifying the thermal response of materials and thus, the thermophysical properties of thermal conductivity (or diffusivity) and specific heat. Many materials yield an anisotropic thermal response, requiring measurements of multiple components of the thermal conductivity tensor. The anisotropy may result from the microstructure of the material (e. g., crystallinity, grain boundaries, molecular alignment in polymers) or the macrostructure (e.g., composite fibers, wood grains). Calendaring of PVC, for example, can produce differences of 30% in the components of thermal diffusivity of an apparently isotropic material (Doss and Wright, 2000). Measurement of the thermal conductivity tensor may also provide a nondestructive test method for the mechanical response of a material if the thermophysical properties can be correlated to deformation (Choy er al., 1978; LeGall, 2000) or stress (Venerus er al., 1999). Several thermo-optical methods are available to measure the components of the thermal diffusivity tensor including thermal wave (Mandelis, 1985), transient grating (Venerus et al., 1999b), and extended flash (Maillet et al., 1990; Doss and Wright, 2000). Each of these methods involves measuring the temperature of a specimen in response to localized heating. Design parameters include the specimen size and shape, the heating area, duration, and magnitude, and the position and number of the temperature sensors, each of which may be a function of the thermophysical properties themselves. Measuring thermophysical properties usually involves estimating two types of parameters: the parameters of interest, such as thermal diffusivity, conductivity, or specific heat, and nuisance parameters, which are parameters that are not of interest to the experimenter but must be estimated in order to arrive at accurate values for the parameters of interest (Beck and Arnold, 1977). Examples of nuisance parameters may include the heat input or convection coefficients. Ideally, experimental design should minimize the influence of errors in either the nuisance parameters or the design parameters of the experiment. The nuisance parameters may be estimated at a reduced accuracy with little effect on the parameters of interest; thus, less computational effort may be expended in finding the nuisance parameters (Beck and Arnold, 1977). Location of the temperature sensors, however, often plays a significant role in the accuracy of the measurements (Emery and Fadale, 1997). Calculating the optimal temperature sensor location, specimen size, or heating geometry requires knowledge of the thermophysical properties to be measured. Of these, only the location of the temperature sensors may be optimized along with the estimated parameters by measuring the temperature field history using, for example, a focal plane array IR camera. The use of such a thermo-optical device allows full-field temperature histories to be recorded, and thus, choice of the sensor locations after the experiment. Recent studies of estimating thermal properties in multiple directions have not fully considered concurrent optimization of the sensor position along with the estimated thermal parameters. Maillet et al. ( 1990) extended the flash method (Parker et al., 1961) to include simultaneous estimates of the radial and axial thermal diffusivities in a cylindrical specimen subject to an infinitesimal heat impulse. They used a geometric method to optimize experimental parameters such as heating radius and eccentric sensor position. Because they used point temperature measurements, they could optimize position only prior to measurements. Doss and Wright (2000) extended the flash method to measure the thermal diffusivity of an anisotropic PVC sheet in three orthogonal directions using an instantaneous heat impulse. LeGall (2001) substituted a finite- duration infrared laser pulse for the instantaneous flash and measured the thermal diffusivity in three orthogonal directions of several elastomers subjected to biaxial finite deformation. Neither Doss and Wright nor LeGall, however, focused on optimizing the experimental parameters or sensor positioning. Instead, the design of their fixed-position thermocouple probe was based on the work done by Maillet et al. and not reoptimized for the geometry or material properties under investigation. Petit and coworkers have focused on replacing detailed numerical models of multidimensional finite flash experiments with lower-order reduced models (Petit and Hatchette, 1998; Le Niliot et al., 2000; Videcoq and Petit, 2001). These studies were primarily analytical and dealt with finding unknown boundary or interface conditions, as opposed to determining thermophysical properties. Graham and coworkers made 2-D thermal diffusivity measurements in orthotropic composite materials using an IR camera to record temperature histories of specimens at multiple locations (Graham et al., 1999; Graham et al., 1999b). Their data reduction was based on a Green’s function formulation that assumes three mutually orthogonal thermal diffusivities and provided simple temperature ratio relations for estimating the diffusivities in two directions simultaneously. While the position of the eccentric temperature sensors were estimated by Graham et al. in computational experiments, the sensor positions themselves were not optimized, and they examined only instantaneous heating of the specimen, which was shown unsuitable for materials of low thermal diffusivity, such as polymers and biological tissue, due to large temperature excursions on the top face and low temperature signals on the bottom face (LeGall and Wright, 2000). Graham et al. also proposed a method for making multidimensional thermal diffusivity measurements in materials with random in-plane orthotropy, analogous to the strain rosette, but reported no such measurements. Emery and Fadale (1997) demonstrated that imprecision in the location of temperature sensors can have a significant effect on the estimated parameters. Sensitivity coefficients (which describe the change in temperature response of a specimen with respect to a given parameter) were computed for an analytical model of a slab experiencing convective boundary conditions on both surfaces. Using the sensitivity coefficients, the determinant of the Fisher Information Matrix, which serves as a measure of the confidence in the estimated parameters, was calculated. Position uncertainty was shown to have the greatest effect on the value of the determinant, and thus, the ability to estimate the thermal properties when sensors were located close to points of high heat flux. 10 Appropriate design of an experiment can reduce the effects of uncertainty in sensor positioning by indicating where points of maximum sensitivity lie via calculation of the sensitivity coefficients and covariance of the estimated parameters over a range of possible experimental designs (Fadale et al., 1995; Emery et al., 2000). Several studies have examined optimal designs for measuring thermophysical properties by analytical means, without experimental comparison. Sawaf and Ozisik (1995) examined measuring orthotropic conductivity in a rectangular parallelepiped by assuming that three sides of the sample are subject to a uniform, constant heat flux while the other three sides are assumed adiabatic. They simulated an experimental temperature response by computing the solution to the diffusion equation using hypothetical properties while adding random noise (on the order of 0 to 1 K, varied between simulations) to the calculated temperatures. They showed reliable convergence to the thermal conductivity used to generate the simulated experimental response using a Levenberg-Marquardt method with an error of less than or equal to 3.2% in each direction. Other recent studies have examined the optimal physical dimensions, heating area, or sensor positions for reliable property estimation. Mejias et al. (1999; 2000; 2003) studied the effects of sensor positioning, heating time, experiment duration, and the use of insulated or constant temperature boundary conditions on the estimation of three thermal conductivity components for a parallelpiped specimen experiencing a constant flux over three surfaces. Dowding and Blackwell (1999) studied the effects of sensor positioning, heating time, heating magnitude, and cooling time on the ability to estimate thermal conductivity and volumetric heat capacity on a l-D model. Aviles-Ramos (2002) studied 11 the sensitivity coefficients developed for the thermal conductivities and heat capacities obtained from a specimen having an orthotropic layer overlaying an isotropic layer. These studies, primarily analytical in nature, often incorporate idealized boundary conditions that may be difficult to achieve physically. For example, Mejias er al. (1999) specified the three heated faces to experience distinct uniform heat fluxes that are functions of time only. Even if the output of the heating device is well characterized, it may be difficult to ascertain precisely the energy that is absorbed by a specimen, or even where that absorption occurs in some materials (see McMasters et al., 1999; LeGall, 2001), and heating multiple faces may lead to additional nuisance parameters. Several studies have investigated the experimental validity of optimal design. The graphical analysis by Maillet et al. (1990) was one of the first attempts to optimize both sensor positioning and heating geometry for multiaxial thermal property measurements. Taktak (1992) investigated sensor position and variation of the heating time and experimental duration on the optimality of measurements of thermal conductivity in composites based on the guarded hot plate method. Taktak used D-optimality wherein the determinant of a matrix of nondimensional sensitivity coefficients of the parameters of interest (i.e., effective conductivity and volumetric heat capacity for Taktak et al., 1993) served as the basis for comparison. Maximizing this determinant results in the minimization of the variation in the parameters of interest. Theoretical results based on l-D geometries were verified using a specimen containing embedded thermocouples. McMasters (1997) considered several different models for the 1-D heating of a specimen that involved departures from the classical flash experiment (i.e., internal radiation, 12 penetration of the flux boundary condition into the specimen, variable conductivity, and a two-sided flash accounting for reflection to the bottom of the specimen surface). Residuals, defined as the difference between the modeled and measured temperatures, were examined to demonstrate the accuracy of the analytical model, while sensitivity coefficients were plotted to illustrate the relative independence of the estimated parameters. McMasters focused on optimizing the data analysis rather than the experimental design. Mandelis and coworkers have developed a thermal wave method of thermal diffusivity measurement that uses the thermal response of the heated surface to determine the thermal characteristics of the specimen. The theory for the thermal wave method has been well described by Mandelis (1985), but the technique may be better suited for detection of cracks or inhomogeneities in materials, rather than measuring thermal diffusivity (Mandelis et al., 1996). Mzali et al. (2004) analyzed 2-D flash diffusivity measurements in which an optimization scheme that maximizes the smallest eigenvalue of the Fisher information matrix was applied to the parameter estimation. The sensitivity coefficients used to construct the information matrix were derived from temperature histories at sensor locations that, in principle, can be relocated during the experiment, say through IR thermography, to optimize the estimated parameters. While IR camera images have been used to determine thermophysical properties in 1-D and 2-D experiments, the results found thus far have only demonstrated that the technique is feasible in 3-D, rather than to optimize the experimental design. Phillipi er al. (1995) used full-field IR images to compute the in-plane diffusivity of an unspecified sample having a diffusivity of approximately 10'6 mz/s. While a sensitivity analysis showed that noise in the temperature measurements would not affect identification of the 13 thermophysical parameters, no attempt was made to optimize the experiment to improve confidence in the results. A complete analysis of the extended flash method should include optimization of physical parameters (e.g., location of temperature sensors), as well as experimental verification of the design of experiment analysis. The research reported here does so by optimizing the measurements using data available from IR thermography and a variable power diode laser. The focus here is on optimizing 3-D thermal diffusivity measurements using D-optimality (Beck and Arnold, 1977) and specifically, optimizing the sensor position for materials having an out-of—plane diffusivity typical of an elastomer and in-plane components varying from one-half to ten times the out-of-plane value. Thermo-optical measurements in atmospheric air are chosen because such boundary conditions are amenable to biological soft tissues and elastomers, which may outgas in a vacuum. The unknown and variable boundary conditions add two nuisance parameters that must be estimated. Experimental demonstration of the optimization technique is made by positioning sensors for a material having an isotropic response typical of a metal, in this case, AISI 304 stainless steel plate. Thus, the goal of this study was to develop a methodology that allows for the optimizing of the temperature measurement positions post-experiment. l4 2.1 Methods The temperature response in a specimen of general anisotropy may be described by the diffusion equation written as [ac-35:— = V(kVT) (2.1) where p is the mass density, c the specific heat, T temperature, t time, and k the thermal conductivity tensor. The thermal conductivity tensor has nine components, in general, and is symmetric (i.e., k3 = k32, etc.) such that only six components are independent (Powers, 2004). The components of k may be considered as being constant over time for small temperature excursions such as those modeled here. When anisotropy is considered, the principal axes of k are often assumed to coincide with the axes used to define the boundary conditions so that only the diagonal terms are nonzero (e.g., Mulholland and Gupta, 1977; Doss and Wright, 2000; Venerus et al., 2001). If the material response is isotropic, the diagonal elements are identical. The thermal diffusivity a, where a = k/pc, is often more convenient to measure than k because a is measured during transient, rather than steady state, temperature fields. Estimating thermophysical properties from temperature measurements requires repeated solution of the diffusion equation while sequentially reducing the difference between the modeled and measured results. Here, an alternating direction implicit (ADI) finite difference solution of equation (2.1) with appropriate boundary conditions is compared to the bottom face temperature history of the specimen as measured at three points. Figure 2.1 illustrates the geometry of a specimen. Specimens are approximately 40 mm by 40 15 20 mm 4\ \ 5.3 mm o \/ A; Figure 2.1 Diagram of the top surface of the slab specimen illustrating the illumination area and the specimen geometry. The quarter plane modeled in the parameter estimation scheme is bounded by the positive x- and y-axes, which themselves are axes of symmetry. Typically, specimens are between 1.5 mm and 3.5 mm thick. Three sensors are located on the bottom surface of the specimen along the coordinate axes a distance d apart; the projections of these sensor positions are shown on the modeled quadrant of the specimen. mm wide and 3 mm thick. The illuminated area covers a 10.6 mm x 10.6 mm square on the top face. Due to symmetry along the x- and y-axes, only the area located inside the positive x- and y-axes is modeled for data reduction. Temperature sensors are located at three positions: one sensor is located at the intersection of the axes at the center of the illuminated area, while the other two locations lay a distance d from the center sensor along the coordinate axes defined by the plane of the specimen. Here, d is to be optimized for a given specimen geometry and composition, and, in principle, can be different for the two in-plane directions. For this simulation, however, the x- and y- 16 diffusivities are assumed equal and this distance is assumed to be the same for both in- plane directions. The analytical problem models an orthotropic planar specimen that has equal in-plane dimensions of length 21 and a thickness 8 that is much smaller than the in-plane half- width 1 (see Figure l). The specimen is heated with a flux qa for a finite pulse time over a square region on the top (2 = 0) surface. Convection acts over the top and bottom surfaces, while the side surfaces see small temperature excursions that may be modeled as adiabatic surfaces. Due to symmetry, only a quarter of the specimen volume shown in Figure 1 need be modeled. The mathematical formulation of this problem may be stated as in - aZT azr azr pC At—k” Az+k22 A’z-t'kx; /azz (2.2) mOSxSLOSySLOSzS&t>0 Initial Condition: T(x, y, z, 0) = T... (2.3a) Boundary Conditions: —k”3§&giAT-Lg=qomz=dnxosxsnosyshbstsn cam h,(T-—T,,°)= k33 BT82 atz=0, forlpr SI, l,,_<_y SLOStStp (2.3c) h, (T - To, ) = k33 3T dz at z = 0, for t > t,, (23d) 3%x=0atx=0andx=l,fort>0 (2-36) a%y=0aty=0andy=l,fort>0 . (23f) 17 hb(T—T°°)=k33aTaz atz=5,fort>0 (2.3g) where k,-,- refers to the axial thermal conductivity in the i-direction, T... the ambient temperature, h, the convection heat transfer coefficient at the top surface, hb the convection heat transfer coefficient at the bottom surface, 1,, the half-width of the heated region, and t,, the heating time. During parameter estimation, the thermal diffusivity components a, (a, = k,, lpc) are estimated instead of the conductivity components kg. The value of hb, the convection heat transfer coefficient at the bottom surface, is assumed known, and is based on results using a 1-D conduction analysis. Temperature histories generated using hb compared favorably to histories generated using a two-layer model that accounted for heat conduction through the air underneath the specimen, however, the computational time was reduced significantly by using hb. The ADI model is unconditionally stable and thus arbitrarily large time steps may be taken during model solution; however, the accuracy of the stable solution must still be verified. Variation of the time steps used here resulted in no change in the computed temperature histories greater than 0.001 K. Studies of grid refinement (in which the grid sizing was changed from 11 x 11 x 11 to 101x 101x 201) showed that a41 x 41x 41 node mesh accurately resolved the temperature history of the specimen to within 0.001 K, and the error in the temperature calculations was within 0.01% based on sz extrapolation. The ADI model operates within a parameter estimation program that was based on a Levenberg- Marquardt iterative nonlinear least squares fitting routine (Press et al., 1992). The parameter estimation program also allowed for spot temperature input from an IR camera and calculated the confidence region of the estimated thermophysical parameters. The 18 Marquardt program uses a x2 merit function to determine the goodness of fit between the data and model. The normalized x2 error is defined as N 2 1 X 2(1); ) = fiZiTj - y 10% i] (2.4) i=1 where N is the number of data points, p,- the parameters being estimated, T,- the data points, and yJ-(pi) the calculated result at the point corresponding to the f“ data point. One criterion commonly used to optimize the design of an experiment, called D- optimality, minimizes the hypervolume of the confidence interval in the estimated parameters as expressed by the covariance matrix (Beck and Arnold, 1977). This design method is well suited for the current analysis, for which the minimization of the covariance matrix is accomplished by maximizing A, defined as the determinant of the Fisher information matrix (Beck and Arnold, 1977; Mejias et al., 1999; Emery, 2001). The Fisher information matrix is composed of normalized sensitivity coefficients Fij (Fij- = pj(dT,-/Bpj), where T, is the temperature measured at position i and p,— is an estimated parameter), which indicate the relative change in the temperature field at a point with a variation in a given estimated parameter. Comparison of different experimental designs requires defining a scaled Fisher information matrix, which may be written as 1. .. 1 .fim. any. ”ill . f... (.5) m,n M'tf s-1 {:0 mapm "apt: mer where M is the number of sensors used (three in this case), If the total time over which data are collected, m and n the indices of the parameters being estimated, T, the 19 temperature at a given sensor and time as calculated by the model using the current estimate for p,-, and Tm, the highest temperature reached during the experiment. It is assumed here that the number of temperature measurements taken in the experiments to be compared is fixed, and that the measurements are sampled at equally spaced intervals during each experiment. Also assumed are the eight statistical assumptions of Beck and Arnold (1977), which may be summarized as: errors are additive and uncorrelated with zero mean and a constant variance, the independent variables are errorless, and there is no prior information available regarding the estimated parameters. Note that the first two parenthetical terms inside the integral are Fsm and F5“, which are calculated as part of the Levenberg-Marquardt estimation scheme, and thus add no penalty to the optimization calculation. Although all of the data obtained in an experiment may be used for the optimization, increasing the information density of the data with respect to the parameters of interest is highly desirable. That is, there may be times with little change in T at some sensors that produce little additional information for determining the diffusivity, for example. Huang (2004) used the ratio of A calculated using all of the estimated parameters divided by A calculated for the nuisance parameters only, A”, as an optimization criterion (Beck and Arnold, 1977). As this ratio increases, the uncertainty in the desired parameters (i.e., thermal diffusivity) decreases. Because the primary concern is reducing uncertainty in the desired parameters, this metric provides a clear basis for comparison of experiments while focusing on the parameters of most importance. Here, optimal sensor positions will be calculated using both A and N to determine whether optimizing the experiment 20 specifically for the parameters of interest has a noticeable affect on the uncertainty in these parameters. To test this procedure for optimizing measurements of the thermal diffusivity tensor, temperature histories were simulated for an extended flash diffusivity experiment for a 40.0 mm x 40.0 mm x 3.35 mm specimen with an out-of—plane diffusivity approximating that of PVC (1.08><10'7 m2/s) and transversely isotopic in-plane diffusivities varying from 0.54x10‘7 mZ/s to 10.8x10'7 m2/s, yielding anisotropy ratios 7 of in-plane to out-of-plane diffusivity of 0.5 < y< 10. Bottom surface temperature histories were calculated at three points: the center of the specimen, and at two points located on orthogonal axes at a distance d from the center point, as illustrated in Figure 2.1. The duration of the simulated experiment and sampling frequency were varied based on the temperature rise of the in-plane temperature sensors. The inter-sensor distance d was varied from 7.0 mm, or slightly outside the projection of the illuminated area, to 20.0 mm, at the edge of the specimen. The inter-sensor position was varied in 1.0 mm increments, except for cases where the maximums of A and A+ lie between neighboring points, in which case the sensor position was moved to the 0.5 mm increment bisecting these two locations to improve interpolation of the optimal sensor location. The heating duration and power were set such that the maximum temperature of the specimen was 100 °C. While this temperature is higher than the glass transition temperature of PVC (T8 = 81 °C, CRC Handbook of Chemistry and Physics, 1991), it is less than the 105 °C regularly experienced by some rigid PVC tubing regularly used to carry electrical cables (Matthews, 1996). Moreover, the temperature excursion in the model exceeds the 80 °C 21 level for no more than 0.4 s for a 1.0 s heating time and 5.8 s for a 10.0 s heating time, and the volume over which T> 80 °C comprises less than 3% of the total specimen volume. Random noise was added to each computed temperature history to simulate measurements such as those acquired with an IR camera. The noise had a Gaussian distribution with zero mean and a standard deviation of o = 0.01 K, which is on the order of the noise resulting from IR thermography. Additional calculations were performed using temperature histories that had added noise with standard deviations of 0.1 K and 1.0 K to determine the effect of the noise level on the data reductions. Once the simulated temperature measurement histories were computed, each served as data for the parameter estimation program, and the three desired parameters (ax, cry, and a2) and two nuisance parameters (the incident heat flux and convection coefficient on the top face) were estimated. The output of the reduction program consisted of the five estimated parameters, the normalized x2 error for the model fit, A and N, and scaled sensitivity coefficients and residual plots for each of the three sensor locations. In order to validate the optimization routine, the optimization was repeated for the sensor positioning on a simulated specimen having thermal properties similar to those of steel, and these results were compared to extended flash diffusivity tests conducted on an A181 304 stainless steel plate. This material was chosen for verification because of its isotropic nature and well-characterized thermal diffusivity (CL = 3.95 x 10'6 m2/s; Incropera and Dewitt, 1990). The 70 mm x 70 mm x 1.50 mm plate was sprayed with a thin layer of flat black stove paint to increase absorption of the laser heat pulse on the top surface of the specimen as well as to decrease the reflectivity and enhance the IR 22 thermographic response on the bottom surface. Four extended flash diffusivity tests were conducted on the plate using heating parameters developed for steel in prior studies (LeGall, 2001). Heating was performed using a 16W diode laser with a beam that was expanded, collimated and then truncated with an aperture plate to achieve the desired heating area. Four tests were performed, each with a heating time of t,, = 1.0 s. Full-field data were sampled at the maximum frame rate of 60 Hz for 30 5. Temperature histories were retrieved from the full-field data at three points on the bottom of the specimen surface: under the center of the illuminated region, and two spots located (I from that center spot, as shown in Figure 2.1. Two values of d were chosen to illustrate the relative ability to estimate parameters at locations both close to and far from the optimal point as determined by the prior optimization analysis. Temperatures for each sensor location were averaged over a 3-pixel by 3-pixel area that was centered over the pixel of interest, in order to reduce random noise. Five parameters were estimated for each test: the three components of the thermal diffusivity tensor (ax, 0ty and a2), a scaled heat flux coefficient, q/pc, and a scaled top surface convection coefficient h/pc. The average values and standard deviations of the out-of—plane diffusivity, a2, and the in-plane diffusivities, 0tx and (1,, were estimated for each sensor position and compared to the values of A and AI calculated from the simulated runs and the measurements to ascertain whether a correlation exists between the accuracy and variation in the values of aand the magnitudes of A and A" for each set of sensor positions. 23 2.2 Results 2.2.1 Measurement Duration Figure 2.2 shows simulated temperature histories and corresponding inverse solutions for four cases of an isotropic material with or = 1.1 ><10'7 m2/s, approximating PVC. Panel (a) shows temperature histories for t,, = 1.0 s and two x- axis sensor positions, d = 7.0 mm and d = 20.0 mm. Panel (b) shows similar temperature histories for to = 10 s and sensor positions of d = 7.0 mm and d = 20.0 mm. One second corresponds to a Fourier number based on the specimen thickness of Foz = (020/52 = 9.62 x 10'3 and that based on the specimen length of Fox = (0t,,t)/l2 = 6.75 x 10“, where 8 is the specimen thickness and l the length of the specimen side. Because the materials modeled in this study are in general transversely isotropic, the y-axis temperature histories are the same as the representative x-axis histories, within the magnitude of the added noise, and are omitted here. Calculations for t,, = 0.1 s and t,, = 10 s were conducted for the isotropic specimen to give insight into the relative magnitude of information derived from each set of tests. The inverse solution fits the simulated experimental results well, as shown by the residuals, defined as the difference between the temperature histories of the simulated experiment and the inverse solution (Figure 2.3). Ideally, the residuals should oscillate about zero with a deviate on the order of the noise exhibited in the temperature measurements. Comparison of A and A+ for different experiments requires each test to be designed so that the same number of data points are collected for each trial (Mejias et al., 1999). Such design allows comparison of experimental design without skewing the results by 24 9 (K) -1 . , . . . 0 50 100 150 200 time (s) -1 . , . . . 50 100 150 200 time (s) Figure 2.2 Simulated temperature histories and inverse model reductions for an isotropic Specimen at sensor locations of d = 0, 7, and 20 mm: (a) 1‘p = 1 s, (b) t,, = 10 s. O is defined as T-T... where To. is the ambient temperature. 25 0.04 0.02 Residual (K) O O O -0.02 -o.o4 . . t . t t 0 50 100 150 200 time (s) 0.04 (b) 0.02 Residual (K) O 8 -0.02 -0.04 1 . I i i O 50 100 150 200 time (s) Figure 2.3 Representative residuals, defined as the difference between the modeled and measured temperature histories, for an isotropic specimen: (a) center temperature sensor, (13) off-axis temperature sensor. 26 using different quantities of data. A common sampling technique to achieve equal numbers of data points is to use the same acquisition rate for the same duration over the experiment. Changing material or experimental protocol, however, may result in these experiments collecting large amounts of relatively unimportant data. The temperature of a thick, low diffusivity specimen, for example, may require hundreds of seconds to reach ambient temperature following a step heat input while a metal specimen may reach ambient temperature after only tens of seconds. Similarly, adjustments in positioning the sensors may yield temperature histories with different times of maximum signal. Consider the temperature response in Figure 2.2a. At the 7.0 mm sensor, the temperature begins to decrease after 65 s, but the temperature at the 20.0 mm sensor has barely risen after 200 5, suggesting that a faster sampling rate and shorter duration are required for the 7 mm sensor, as compared with the 20 mm sensor, to retrieve similar amounts of useful information. Instead of fixing the sampling rate and duration, it may be advantageous to fix the number of data points sampled during the temperature transient and vary the sampling rate and duration based on some other criterion. A series of computational experiments was run for an isotropic specimen to investigate the quality of information taken from such a test. Each test contained 2000 data points at each of three sensors over the duration of the experiment (for a total of 6000 points). The sampling rate and duration were varied to yield 6000 points from a minimum duration of about 20 s to a maximum of 250 5, thus capturing the temperature histories from the initial temperature rise on the bottom center of the specimen to the cooling of the bottom specimen surface. Three inter-sensor distances were chosen to give insight into the effects of changing the sensor location on the optimum sampling duration. Figure 2.4a shows the in-plane 27 temperature response of an isotropic specimen for d = 12 mm and the corresponding sensitivity coefficients. Figure 2.4b shows the values of A and A+ as a function of elapsed time for the same sensor location. Note that the maximum values of A and A+ occur at approximately the same time as the maximum value of the scaled heat flux sensitivity coefficient qu, as shown in Figure 2.5. For tests conducted using small d, these maxima occur slightly after the in-plane sensors reach their maximum temperature, but as d increases, the maximum of the in-plane temperature response occurs at approximately the same time at which A, N, and qu reach their maximum values. These results suggest that sampling the temperature histories until each of the in-plane temperature sensors has reached its maximum value should yield the most robust results for each test. Extended flash diffusivity measurements on PVC and steel by Doss (1998) and LeGall (2001) corroborated these results, as the variability in the estimated diffusivities was minimized when the measurements were conducted until maximum rise of the in-plane temperature sensors. 2.2.2 Anisotropic Diffusivity This approach has been extended to materials with an orthotropic thermal response. Simulations without added noise were used to identify the time of the maximum in-plane temperature excess for 0.5 S y S 10 and 6 mm S d S 20 mm. Once these times were identified, data were simulated again by adding noise to temperatures calculated using the parameters noted above. The sampling rate and duration were adjusted to yield 6000 data points (2000 points from each of the three sensor locations) over the calculated time of maximum temperature. These data were then used as input for the parameter estimation scheme, which again estimated five 28 1.0+ (b) 2‘ 0 OD OD O 0.0- u‘ 0 50 100 150 200 250 time (s) Figure 2.4 Panel (a) shows the in-plane temperature response and Hg for an isotropic specimen where d = 12 mm. The dots represent the in-plane temperature data, the solid line the CL, sensitivity coefficient, the dotted line the heat flux sensitivity coefficient, the short dashed line the (1.1 sensitivity coefficient, the long dashed line the or, sensitivity coefficient, and the dot-dash line the convection sensitivity coefficient. Panel (b) shows A (circles) and A” (triangles) as a function of elapsed time for the same specimen and sensor location. 29 -7 ' . 4 ' ' 0‘ t ' 00 .v “ J .a - -.. 0. fl 1.! ‘ ¢ . ’v‘. D. l A V J "\.- ' '. . v ' . I'I - . O o ' - . 1 ' . -' . -.I . .‘ - p. n, t "' .. .' .. -~.--.~ ~.« I . y— n - p . 5 ‘. - a, - ,‘ v~\. I fit . - o _-.- 0 f . ' w ... Q ‘ L. . o ! ' \ 4 . 't ‘ ' I ‘ .,_, .J c I <_“ n" .' P '.‘«:‘< .“ '3" "3: 9 — ‘Kfi. " p— o 2 - - __ . . o . log A+ 9 (K) -1 0 ‘ 39:" “ 0.1 -11 4E7 ' 0-0 -12 ‘ I ' l ' l ’ ' T '01 0 50 100 150 200 250 time (s) Figure 2.5 In-plane temperature response (circles) and log A+ (triangles) as functions of time for an isotropic specimen with d = 12 mm. 30 parameters and calculated A and A“. Figure 2.6 illustrates the typical variation of A and A+ for changing d, with the maximum for each A and A+ lying at d = 8 mm. For each y, the out-of—plane diffusivity and the average of the two in-plane diffusivities were resolved to within less than 1%. Figure 2.7 shows how the optimal d varies with y as described by the maximum A and N. These maxima occur for sensor positions near the illuminated region for small 7, and closer to the edge of the specimen as 7 increases. Note also that the maximum values of A and A” occur at the same location for y < 6, but separate for larger values of 7. For 7 = 9 and 10, the maximum values of A and A+ occur at d = 20 mm, the edge of the specimen, suggesting that the specimen width must be increased to accommodate materials with y > 8. Due to the ambiguity in the meaning of these results, only results for y S 8 will be discussed further. 2.2.3 Experimental Validation Extended flash diffusivity tests were conducted on A181 304 stainless steel to demonstrate the accuracy gained in the estimated parameters when using optimized sensor positioning. The analytical optimization routine detailed previously was repeated for a simulated 1.50 mm specimen having an isotropic thermal diffusivity (or = 3.95 x 10'6 mzls) and heat flux and convection boundary conditions similar to those in prior flash diffusivity tests (LeGall, 2001). As shown in Figure 2.8, the simulated experiments indicated that the optimal sensor position as quantified by both A and A+ should be as far from the illuminated region as possible, suggesting that d should be set to 20 mm for a 40 mm x 40 mm specimen. In principle, such an experiment might yield the results with highest confidence, except that the assumption of the specimen edge being adiabatic is violated, as there is a small temperature gradient at 31 0 o . o -4* . ‘ --8 ‘ o o <’ “<1 8’51 ‘ ”-93 A o '5‘ . --10 A o o '7 I'l'I'TTFTI'I'I‘1*I'I'I'I*T‘-11 6 7 8 9101112131415161718192021 d(mm) Figure 2.6 Variation of A and A+ with respect to d for an isotropic specimen with 1,, = 1.0 8. Calculated values of A are represented by circles, while calculated values of A“ are represented by triangles. 32 21 19‘ 17 ‘ 15‘ Figure 2.7 Location of the maximum A (circles) and A" (triangles) as a function of y. the specimen edge. Additionally, prior measurements with the IR camera suggest that the temperature rise at the off-axes sensors is on the order of the experimental noise for d > 15 mm and a heating time 1,, = 1.3 s. In light of this, sensor positioning was limited to the area bounded on the interior by the shadow of the heated area and the exterior by roughly the outer third of the specimen half-width and, thus, 6 mm < d < 14 mm; specifically, d = 8.45 mm and d = 13.75 mm were chosen for the IR measurements. These values of d correspond to distances of 40 and 65 pixels, respectively, from the central illuminated pixel in the IR camera field of view. Results of these measurements are shown in Table 2.1, Table 2.2 and Table 2.3. Note that the average of the estimated 0t" and ory using d = 33 8.45 mm differs from the reference value by almost 5%, but the difference between the reference and average estimated in-plane value using (I = 13.75 mm drops to 0.5%. The error in the estimates of 0tz are higher in magnitude, over 11% for d = 8.45 mm, but about 5% for d = 13.75 mm. Table 2.3 lists the experimentally obtained values of A and AJ, for the steel plate tests, which are, in general, lower than the values calculated from the analytical simulations. The estimated values of the nuisance parameters (heat flux and convection coefficient) were slightly lower than those used for the simulation, resulting in lower calculated values of Fij, A and N. For tests conducted at the same d, the average values of A and A+ are different for different heating durations. A longer heating duration creates higher temperature excursions in the specimen and, therefore, larger scaled sensitivity coefficients, increasing the value of A. Note that the average values of both A and A+ are about three to four times larger for the d = 13.75 mm tests, where the estimated diffusivity values are closer to the reference values, as compared to those values obtained for d = 8.45 mm. 34 '4 -8.5 , '5_ ‘ A . A . O . ' --9.3 -6“ ‘ . A ‘ . . <1 '7 . +<, U) -- 2 A . 10.0 .8) -8“ C .9n ”-10.8 40* ' '11'1'I'I'r'l'i'r*r't'1't'tvrtrv-11.5 6 7 8 9 1O 11 12 13 14 15 16 17 18 19 20 21 d(mm) Figure 2.8 Variation of A and A+ with respect to d for an isotropic steel specimen with t,, = 1.0 5. Calculated values of A are represented by circles, while calculated values of A+ are represented by triangles. 35 Table 2.1. Estimated directional thermal diffusivity values and associated standard deviations for a 1.50 mm stainless steel plate (average of four tests). (I avera e orz on average (itx ow, average 0ty on, (mm) (mm ls) (mmzls) (mm2/s) (mm2/s) (mmZ/s) (mm ls) 8.45 l 3.50 ‘ 0.13 l 3.74 i 0.07 , 3.78 l 0.05 13.75 3.75 0.14 4.01 0.06 3.93 0.04 Table 2.2. Average deviation of the out-of-plane and in-plane diffusivities from the reference value of or = 3.95 x 10’6 mzls for a 1.50 mm stainless steel plate. 61 average or, deviation average 01,, 0ty deviation from (mm) from reference (%) reference (%) 8.45 I -11.4 ’ —4.8 13.75 -5.1 0.5 Table 2.3. Values of A and A+ computed from measurements of the 1.50 mm steel plate d (mm) I average A I average A+ 8.45 3.02 x 10‘" 2.43 x 10‘l2 13.75 1.30 x 10'9 8.03 x 10'12 36 2.3 Discussion The estimated values of thermal diffusivities in the analytical portion of the study all fell within 0.2% of the input diffusivities for 1 S 7S 8 with added noise having a standard deviation of 0.01 K. The uncertainty in the estimated diffusivities reported in the literature for multiaxial measurements is approximately an order of magnitude larger than the uncertainty calculated here. For example, measurements made by Doss and Wright (2000) on A181 304 stainless steel exhibited a standard deviation of 3.5% of the reference diffusivity value in the out-of-plane direction and an uncertainty of 5.6% over the in- plane directions, while 2-D measurements on PVC by Maillet er al. (1990) showed a variability in the estimated diffusivity value of 3% between tests in the axial (out-of- plane) direction and 5% in the radial (in-plane) direction. Doss and Wright reported noise on the order of 0.1 K, which may result in the greater uncertainty. The higher accuracy of the computational experiments may be expected, because the same analytical model that generated the temperature histories, after addition of noise, was also used in the parameter estimation, and thus the simulated experimental boundary conditions are perfectly modeled. Moreover, the noise in the simulations conforms to the eight standard assumptions for parameter estimation (Beck and Arnold, 1977). For 7 = 0.5, accuracy decreased dramatically for d > 13 mm, with the error in the estimated values of the thermal diffusivities exceeding 4% or more for the out-of—plane direction and 70% for the in-plane directions. For these cases, the in-plane temperature rise was less than 0.07 K, leading to low signal-to-noise ratios and sensitivity coefficients of low magnitude. The 37 lack of a strong signal results in poor quality information at those sensor locations and thus, the inability to estimate the parameters properly from such data. The information obtained from each experimental simulation, as quantified by A for the optimal sensor position and heating time, varied from a maximum of 1.35x10'3 for y = 0.5 to 1.00><10'6 for y = 8. For A”, a maximum value of 1.58x10'8 was calculated for y = 1 and a minimum value of 2.83><10"0 for y: 8. Table 2.4 shows the maximum values of A and N along with the corresponding (1 as a function of 7. Note that both optimality criterion yield identical d for 1 < y < 2, a range typically seen in deformed or machined elastomers (e.g., Doss and Wright, 2000; Venerus et al., 2001). One might reason that the closer the sensors are to the heated area, the higher the temperature signal and, presumably, the more information is obtained from the experiment. Results based on A and A’( indicate that this is not necessarily the case, however, as the optimal sensor position may be located far from the illuminated area, as shown in Figure 2.7 for elastomeric specimens having a large 7. For 7 < 6, the optimum sensor position lies at about the same location, either 8 or 9 mm, for either criterion. For 6 < y < 8, however, the optimal position moves further from the center, and this location differs depending on whether one bases optimization on the maximum A or A+ (see Table 2.4). While prior studies have looked at the optimization of sensor positioning (e.g., Maillet et al., 1990; Mejias et al., 1999; Mzali et al., 2004), the optimal position was not examined as a function of anisotropy ratio as it is here. Optimizing the measurements as a function of varying anisotropy is important because the thermophysical properties of elastomers have been shown to vary with mechanical loading (e.g. Broerman et al., 1999). Since 38 Table 2.4 Optimum sensor location based on A and A“ for various y in a phantom elastomer with an out-of—plane diffusivity of l.08><10'7 m2/s and 1,, = 1.0 s. y optimum A d for optimum A optimum A+ d for optimum A+ (mm) (mm) 0.5 1.35 x 10'3 8 1.32 x 10'8 7 1.0 7.56 x 10“ 8 1.58 x 10'8 8 2.0 1.49 x 10“ 9 7.63 x 109 9 3.0 5.00 x 10'5 9 4.50 x 109 9 4.0 1.56 x 105 9 2.14 x 10° 9 5.0 6.38 x 10*5 10 1.27 x 109 10 6.0 2.79 x 10'6 12 6.78 x 10'10 10 7.0 1.43 x 1045 13 4.19 x 10'10 11 8.0 1.00 x 106 16 2.83 x 10'l0 12 material thermophysical properties may be unknown prior to estimation, especially if the material is deformable and affected by loading during the experiment, repositioning of the experimental temperature sensors based on the measured properties is necessary to yield the most accurate results. Methods of temperature measurement that allow flexibility in sensor positioning post-data acquisition enable an iterative data analysis that will yield better estimates of thermal pr0perties than those afforded by fixed-position methods of temperature measurement, such as thermocouple probes. Variation of the simulated heating time seemed to have little effect on the location of the optimum sensor position. Simulations using 1,, = 0.1 s were inconclusive, as the simulations for d > 1 1.0 mm did not converge for that heating time due to the low temperature signal resulting from the imposed 100 °C temperature limit on the top face of the simulated specimen. Table 2.5 shows the relationship between t,, and the maximum values of A and A”, as well as the maximum error in the estimated diffusivity values for each heating time. While the information available to the reduction program for 1,, = 0.1 s 39 was enough to resolve the diffusivities to within 1% for d S 11 mm, the differences in magnitude of A and A” near the optimum were very small (less than 3%), leaving the optimal d unclear, although it appeared to be between 8 mm and 9 mm. Results for 1,, = 10.0 5 indicated an optimal sensor position of 8.0 mm based on A”, similar to the 1.0 s heating time tests, but the optimal sensor position using A shifted to 9.0 mm. As the heating time increased, the optimum values of A and A” increased by orders of magnitude, suggesting that the accuracy of the estimated parameters would improve with increasing heating time. Prior measurements have shown that extended heating (e.g., FoZ > 0.2 corresponding here to t> 20 5), even at low power, can complicate the modeling of the convection on the specimen surface, as well as cause large temperature gradients near the specimen edges, both of which depart from the modeled behavior (data not shown). As a result, simulations having longer heating times were not investigated, despite A and Table 2.5 Relation between t,,, optimum A and N, and variation in or. ’1) (S) optimum A optimum A+ maximum error in 0t (%) 0.1 3.31 x 10‘8 2.73 x 10'” 0.5 1.0 7.56 x 10“ 1.58 x 10'8 0.1 10.0 6.50 x 100 1.43 x 10‘5 < 0.1 A+ both increasing as 1,, increased. Note that the maximum variation in the estimated diffusivity values in Table 2.5 lies below one percent, even for the cases with short heating times. Given the accuracy possible with the 1,, = 0.1 5 case, one might conclude that the shortest heating time that gives a clear signal may be acceptable for running such transient thermophysical property measurements. Experiments using essentially instantaneous flash times (Degiovanni, 1977) may require such a large heating power (to 40 obtain a good temperature signal at all sensor positions) that the top face of some specimens may ablate or undergo a phase change, which would be undesirable. Thus, if the total energy input to a specimen is to be held constant, the longest possible heating time should be used that minimizes edge effects or complicated convection boundary conditions. The level of noise present in the isotropic data affected the accuracy of the estimates for the diffusivities, but did not affect the optimum value of d. Near the optimum d (:2 mm from the optimum location), the accuracy of the estimated diffusivities was approximately the same for both the out-of-plane and in-plane directions, but the error in the estimated out-of—plane diffusivity was approximately three times lower than the associated error in the in-plane directions when the sensor position was far from the optimum d (see Table 2.6). Note that, for the case of maximum noise, when 0' was equal to 1.0 K, the variation in the out-of-plane and in-plane diffusivities was at most i10.6% when d was close to the optimum, but this variation increased to i27.7% for the in-plane diffusivity far from the optimum configuration. While the magnitude of A at the optimum location decreased by 47% as the standard deviation of the noise increased from 0.01K to 1.0K, the magnitude of N at the same points remained nearly constant. Because AI relates the ratio of the overall information divided by the nuisance information, a decrease in the overall information A, with A” remaining constant indicates that an increase in noise level has a negative effect primarily on the quality of the information in the nuisance parameters. While attempts should always be made to minimize noise and improve the signal-to-noise ratio when conducting flash diffusivity experiments, these 41 Table 2.6 Relationship between noise level in measurements and the corresponding accuracy in the thermal properties of interest for an isotropic specimen with t,, = 1.0 s. 0' (K) maximum variation for (itz maximum variation for 01., far optimum value near optimum d (%) from optimum d (%) of A 0.01 < 0.1 0.1 7.56 x 10'4 0.1 0.5 0.5 7.00 x 10“ 1.0 4.8 6.1 3.97 x 104 c (K) maximum variation for maximum variation for Other, optimum value abut, near optimum d (%) far from optimum d (%) of A+ 0.01 < 0.1 0.5 1.58 x 10'8 0.1 0.6 3.5 1.59 x 108 1.0 10.6 27.7 1.58 x 10’8 results indicate that extreme efforts to reduce noise in the experiments may not result in an improvement in the accuracy of the estimates of the thermal diffusivities, since the noise primarily affects the estimation of the nuisance parameters. However, the use of IR thermography to measure temperature has reduced the level of noise by an order of magnitude as compared to E-type thermocouples for measurements in our laboratory, and this noise reduction corresponded to a reduction in the number of iterations needed to converge on the estimated parameters. Thus, the benefits of noise reduction may extend beyond the accuracy of the parameters. While the analytical results for y < 6 suggested that either A or A” could be equally useful in determining the optimal d, the reduction in computing time brought on by the improved signal-to-noise ratio suggest that the relative magnitude of experimental noise affects the certainty in the parameters of interest via a decrease of information in the nuisance parameters, and as such, A may be more useful than A+ for experimental optimization. 42 Experimental validation of the optimization procedure was performed using a stainless steel plate. Analytical results showed that more information was obtained about the parameters of interest (as quantified by A and A‘“) at the far temperature sensor location despite the overall lower temperature signal at that position. The increase in information obtained at the optimal sensor position resulted in estimated diffusivity values that were closer to the reference diffusivity of A151 304 stainless steel than diffusivities estimated using the lower information available at the non-optimal sensor location. The standard deviation of the estimated in-plane diffusivities compared favorably to earlier experimental results by Doss and Wright (1.3% of the reference value for these measurements versus 5.6% for Doss and Wright), while the standard deviation in the out- of-plane diffusivity 0tz was unchanged (3.5% in both sets of experiments). However, the estimated value of (1,, reported here was 5.1% lower than the reference value, while the value reported by Doss and Wright was only 1.8% lower than the reference value. This discrepancy in 0t; is likely due to the 60 Hz frame rate of the IR camera. The camera frame rate is constant for a given test and does not synchronize to accuracy greater than 60 Hz with the heat pulse trigger, and as a result there is some scatter in the estimated out-of-plane diffusivities over successive tests. The temperature response at the off-axis sensor locations is slower, and as such, a temporal error amounting to a fraction of a time step does not affect the estimated in-plane diffusivity values. This temporal effect is magnified for materials of high thermal diffusivity, such as steel. When the steel data temperature histories were adjusted by adding or subtracting a single time step of 1/60 s, the estimated value of orz varied by 17.5%. When similar adjustments were made to the temperature history of a material having a smaller thermal diffusivity (PVC sheet; at z 1.3 43 x 10’7 mZ/s), the estimated value of 0tz varied by less than 1% despite the use of a slower frame rate of 15 Hz to capture the PVC data. Because this timing error is random, it is expected that the average value of (1tz will approach the reference value as the number of measurements in the sample is increased. Still, the results gained from the optimal sensor location reduced uncertainty in the experimentally derived values of diffusivity, and as such, shows that this method holds promise for improving future extended flash diffusivity measurements. 2.4 Closure Estimation of the axial components of the diffusivity tensor was possible to within a 0.2% error for all experiments featuring a 1.0 s heating time, an CL, = 1.08x10'7 m2/s, and l < "y < 10. The optimum d was determined via two D-optimality parameters: A, which gives a measure of the total information available from an experiment, and A+ which is the ratio of A as defined above divided by A for just the nuisance parameters (here, heat flux and convection). This distance d was shown to be a function of y for the analytical experiments simulated here. Increasing the pulse time results in both a higher level of total information and a higher level of information concerning the parameters of interest, but physical complications to the experiment may outweigh the slight improvement that results in the accuracy of the estimated parameters. Experimental measurements made on a stainless steel plate showed that measurements made closer to the optimal sensor positions as suggested by this analysis lead to estimated parameters that are closer to the expected reference values. These results support the use of IR thermography, which allows for more flexible sensor positioning and lower measurement noise than other fixed-position methods of measurement for transient extended flash thermal diffusivity measurements. Overall, with this method showing promise in optimizing experiments on specimens having transverse isotropy, a next step would be to investigate materials possessing thermal orthotropy. 45 CHAPTER 3 ON MEASURING THE THERMAL DIFFUSIVITY TENSOR WHEN PRINCIPAL AXES ARE UNKNOWN A PRIOR! 3.0 Motivation When modeling the anisotropic thermal response of materials, the material may be rotated such that the physical axes coincide with the principal axes of the thermal diffusivity tensor, resulting in thermal orthotropy. This may be impractical for a specimen of arbitrary shape. During measurements of such a tensor, however, the principal axes may be unknown, requiring a method to determine principal values and the orientation of the principal directions while simultaneously measuring the diffusivity. Thermal anisotropy may result from the microstructure of the material (e.g., crystallinity, grain boundaries, molecular alignment in polymers) or the macrostructure (e.g., composite fibers, wood grains). Deformation and manufacturing may induce anisotropy. Calendaring of PVC, for example, can produce differences of 30% in the components of thermal diffusivity of an apparently isotropic material (Doss and Wright, 2000). Measuring the thermal diffusivity tensor may thus provide a nondestructive test method for the material response to a mechanical state if the thermophysical properties can be correlated to deformation (LeGall, 2000; Choy et al., 1978) or stress (Venerus et al., 1999). 46 The value of each component of the thermal diffusivity tensor depends on the reference orientation from which the specimen is viewed. The temperature response in a specimen of general anisotropy may be described by the diffusion equation, written as pc%§=V(kVT) (3.1) where p is the mass density, c the specific heat, T temperature, t time, and k the second- order thermal conductivity tensor. The thermal conductivity tensor, in general, has nine components: Fkr 1 kn k13 1 k= [<21 1‘22 k23 (3.2) H31 1‘32 1‘33 k is a symmetric tensor with only six independent components, as required by Onsager’s theory of reciprocity, and thus kij = kji (Onsager, 1931; Carslaw and Jaeger, 1959; Powers, 2004). For small temperature excursions, the elements of k may be considered constant over time. The thermal diffusivity a, where a = k/pc, is often more convenient to measure than k because a is measured during transient, rather than steady state, temperature fields. The diagonal terms of 11(01“, an, (133) are associated with axial conduction in Cartesian coordinates, while the off-diagonal terms, here called the shearing diffusivities, define heat conduction along the lines that bisect the axes indicated in the mixed derivative dleaiaj associated with each cg,- term. The effect of the shearing diffusivity acting in the 1-2 plane is illustrated in Figure 3.1. Figure 3.1a shows temperature contours on the bottom surface of a specimen having equal axial diffusivities in the 1- and 2-directions with no accompanying shearing diffusivity, resulting in transverse isotropy, while Figure 3.1b shows the same contours for a specimen having the 47 same axial diffusivities as before but now also possessing a shearing diffusivity that is about 18% of the axial diffusivity value. Note how the addition of the shearing diffusivity causes elliptical-shaped temperature contours with major and minor axes oriented towards the lines y = x and y = -x, as opposed to the circular contours that would result from transverse isotropy (as shown in Figure 3.1a). The change in isotherm shape due to the effects of the shearing diffusivity is analogous to the distortion experienced by a material element that is deformed under a shearing strain in solid mechanics, hence the term “shearing” diffusivity. If the principal axes of thermal diffusivity coincide with the physical axes of the experiment, the off-diagonal shearing terms have zero magnitude, resulting in material orthotropy (Figure 3.1c). Rotation of the case presented in Figure 3.1b clockwise by 45° results in the same contours shown in Figure 3.1c. The temperature response depicted in Figure 3.1b may result from an orthotropic material having principal axes that are rotated an angle B from the physical axes of the experiment. Several studies have presented analytical models of orthotropic materials useful for thermophysical property estimation. Mulholland and Gupta (1977) discuss the 3-D solution of the diffusion equation for an anisotropic solid with boundary conditions that are difficult to solve in an orthonormal coordinate frame. The model is rotated such that the principal and physical axes coincide and then the coordinate lengths are scaled by a factor that renders the resulting equation isotropic. The transformation of an orthotropic material response into an isotropic response via appropriate scaling of the coordinate axes is discussed further by Ozisik (1993). Chang and Tsou (1977a, 1977b) developed an 48 20 20 (a) , (b) 10‘ 10* E ’5‘ e 0* e 0: > > -1o< -10‘ -20 . . . _20 . . r -20 -10 O 10 20 -20 -1O 0 1O 20 X (mm) 20 x(mm) (C) 10‘ ’e‘ e 0‘ > -10' -20 . . . -20 -10 0 10 20 x(mm) Figure 3.1 Isotherms on the bottom surface of a specimen at t = 30 8 following 1 s heating. (a) transversely isotropic: or” = 0122 = 0.16 mmzls, 01.2 = 0, (b) anisotropic: 01“ = 0122 = 0.16 mmz/s, (1.12 = 0.0282 mmzls, (c) orthotropic: or“ = 0.188 mmzls, 022 = 0.132 mmzls, 002 = 0. Note that Figure 3.1c is congruent to Figure 3.1b if rotated 45° counterclockwise. algebraic transformation using conformal mapping that reduces a thermally anisotropic problem to an isotropic one, but this technique was restricted to problems formulated in cylindrical coordinates having homogeneous boundary conditions. Poon et al. (1979) extended the conformal mapping technique to specific 2-D and 3-D cases in both 49 rectangular and cylindrical coordinates with homogeneous boundary conditions. Huang and Chang (1984) extended this conformal mapping technique to 2-D geometries having mixed boundary conditions (i.e., flux at one boundary, fixed temperature at another). Haji-Sheikh and Beck (2002) developed an analytical solution for the temperature field in 3-D, multi-layer bodies where the principal thermal and physical axes coincide, leading to thermal orthotropy, and thus, three nonzero terms in the 01 tensor for each material layer. Aviles-Ramos (2002) and Mejias et al. (1999; 2000; 2003) have investigated the optimization of sensor position and experimental boundary conditions for orthotropic specimens. Pron and Bissieux (2004) have developed a 3-D model of heat diffusion in orthotropic media for investigation of stress-induced orthotropy. Mzali et al. (2004) analytically investigated the effect of sensor positioning on the estimated diffusivities involving an orthotropic cylindrical specimen. Recent analytical treatments of property measurements on anisotropic thin films (Bennett, 2004) and a 2-D cylindrical slab (Milosevic and Raynaud, 2004) have assumed that there is no shearing conductivity that acts between the axial and radial directions, and furthermore, that the radial conductivity is isotropic with respect to in-plane orientation. Powers (2004) lists necessary conditions on the conductivity tensor that allow only physically meaningful anisotropic problems; namely, the conductivity (or diffusivity) must be positive semi-definite and that Onsager’s reciprocity theory must hold (00,- = 09,). Aside from a treatment of a steady- state 2-D solution for slab-like multilayer samples (Hsieh and Ma, 2002; Ma and Chang, 2004), little work exists in the recent literature that investigates the response of materials oriented in other than the principal directions, which yields a fully populated diffusivity tensor. 50 Prior experimental techniques used to measure the thermal diffusivity tensor have assumed that the material response is orthotropic, resulting in nonzero values for only the diagonal terms in 0. Early extensions of the flash method that looked at both an axial and radial component of thermal diffusivity assumed no coupling between these two terms (Donaldson and Taylor, 1975; Chu et al., 1980) and thus no shearing diffusivity need be considered. Doss and Wright (2000) extended the 1-D flash method of Parker et al. (1961) to three directions through estimation of the two principal in-plane components of or via the temperature history of a specimen at three distinct points. It was assumed that the material either exhibited no shearing diffusivity or that the principal axes of a were aligned with the physical axes of the experiment, resulting in thermal orthotropy. Graham et al. (1999, 1999b) suggested an experimental format to measure four of the six independent components of the thermal diffusivity tensor in a slab specimen. Their formulation assumed that the out-of-plane direction is a principal direction, and thus there were no shearing terms involving the out-of—plane axis. The in-plane orientation was assumed unknown, and it was suggested that ratios of temperature histories obtained from the top and bottom surface of an anisotropic specimen may be used to ascertain both the three principal diffusivities and the angle B that the principal in-plane axes make with the physical axes of the specimen. No data were presented, however, to illustrate the use of this experimental design for thermal property estimation. Also, while Graham et al. did estimate the sensor location in each analytical experiment, the experimental sensor positioning itself was not optimized. Because the assumption of thermal orthotropy seems prevalent in recent experimental studies, quantification of the error inherent in 51 assuming orthotropy for parameter estimation studies in a material exhibiting randomly oriented anisotropy would be of value. Such a study could establish conditions for which orthotropy could be assumed in a material exhibiting a fully populated 0t tensor. To investigate the effects of randomly oriented anisotropy on the temperature response of a specimen, computational models need to be written that account for the thermally shearing terms. Orthotropic models are easily implemented in an implicit fashion using an ADI formulation of the finite difference method (e.g., Douglas, 1962) that offers substantially quicker computations than full-matrix inversions, however, full-matrix inversions must be performed when modeling any shearing diffusivities using an implicit finite difference scheme. The shearing terms in the a tensor are not readily implemented in an ADI formulation, for example, since this method uses operator splitting to divide each time step into a series of substeps in which only one direction at a time is treated implicitly (see Press et al., 1992 and Pozrikidis, 1998). It may be easier to implement finite difference models of the diffusion equation in an explicit formulation when shearing terms are present in the a tensor. The stability criterion for the explicit problem is not affected by the shearing diffusivities, as the contribution of these terms negate each other. The runtime for such a problem would be on the order of twice the explicit formulation of the orthotropic problem, since modeling of the fully populated 01 tensor involves computation of three shearing terms on top of the three axial terms already modeled in the solution of the orthotropic heat diffusion equation. 52 To speed modeling of the thermal response of a randomly oriented anisotropic specimen, one of the physical axes of the problem, here the out-of-plane axis for a slab specimen, may be assumed to coincide with a principal thermal direction of the specimen. In such a case, four of the six off-axis terms would be zero, and the a tensor may be represented as 0‘11 CL12 0 a: 0121 0I22 0 (3.3) _ O O (133_ with 002 = (121 due to material symmetry (Powers, 2004). Physically, this tensor would represent the thermal behavior of an orthotropic specimen with principal diffusivities of an unknown orientation in the 1-2 plane. Given such behavior, the parameter estimation routine could be used to seek a solution based on the physical axes imposed here, leading to the four estimated parameters (11 1. 0122, 0112, and 0133, or, a solution could be sought based on the principal axes of the specimen, requiring estimation of four different parameters: the principal thermal diffusivities of the specimen, 01,, 0t", and 01111 (which here, by definition, is equal to 0133), and B, the angle between the principal and physical axes in the 1-2 plane. The assumption of one physical axis aligning with a principal direction is a reasonable one; as for an example, an elastomer rolled into thin plates may exhibit such behavior, as the manufacturing process for such plates may be shown to impose anisotropy in the plane of the specimen while leaving the out-of-plane diffusivity unaffected (Doss and Wright, 2000). As another example, the physical geometry of lamellar composites favors such an assumption, since the orientation of the lamellae typically does not favor reorientation of the principal axial diffusivity away from the out- of-plane axis. Assuming the suggested thermal orientation, the goal would be to design 53 an optimal experiment to resolve the non-zero components in the a tensor without knowledge of the angle needed to rotate the specimen into the principal orientation and thus be able to make better measurements of the given a tensor in materials without prior knowledge of specimen orientation. Optimization of thermophysical property measurements becomes increasingly challenging as the complexity of the experiment increases. While 1-D thermal property experiments such as the flash method (Parker et al., 1961) or guarded hot plate method (e.g., Bradley, 1943) are simple in design and therefore relatively straightforward to optimize, experiments used to estimate thermal properties in two or more directions require well defined heating protocols, boundary conditions and sensor positioning and thus a more involved optimization analyses. Several thermo-optical methods available to measure the directional components of the thermal diffusivity tensor include the thermal wave (Mandelis, 1985), transient grating (Venerus et al., 1999b), and extended flash methods (Maillet et al., 1990; Doss and Wright, 2000). Each of these methods involves measuring the response of a specimen to localized heating and each should be designed to provide the greatest certainty in the property values reported. Design parameters include the specimen size and shape, the heating area, duration, and magnitude, and the position and number of the temperature sensors. Prior work has focused on the optimization of sensor position and heating power and duration for thermal diffusivity measurements on an orthotropic slab specimen (Davis and Wright, 2005). There, three parameters of interest (or, 1, 0122, 033) and two nuisance parameters, defined as parameters that must be estimated along with the parameters of interest in order to solve the inverse problem, but 54 are not valued for themselves (Beck and Arnold, 1977), in this case, incident heat flux and top face convection, were estimated. Here, the temperature fields are calculated, so what might be nuisance parameters are assumed known, leaving only the four non-zero components of the diffusivity tensor to be estimated. The specimen geometry is held fixed, as are the sensor locations and heating level and duration, which were all based on results from the prior study. The current investigation focuses on ( 1) an investigation of the nature of the sensitivity coefficients resulting from a material exhibiting in-plane anisotropy and random orientation, (2) quantification of the error inherent in assuming orthotropic behavior for a specimen whose principal thermal axes are rotated an arbitrary angle from the physical axes, and (3) optimization of the measurement duration given a specimen of random in-plane anisotropy. 55 3.1 Methods Estimating thermophysical properties requires repeated solution of the diffusion equation with the goal of sequentially reducing the difference between the model results and measured temperatures. For a simulated experiment, data are first generated and substituted for measurements as input for the inverse problem. Here, an explicit finite difference model solution is compared to the bottom face temperature history of a hypothetical specimen as measured at four points. The analytical problem models an orthotropic planar specimen that has equal in-plane dimensions of length 21 and a thickness 5 that is much smaller than the in-plane half-width l. The specimen is heated with a flux go for a finite pulse time over a circular region on the top (2 = 0) surface. Convection acts over the top and bottom surfaces, while the side surfaces see small temperature excursions that may be modeled as adiabatic surfaces. The mathematical formulation of this problem may be stated as 3T =k 32y k BZT 2k 32y +k 32y 3.4 '06 At 11 ax2+ 22 ay2+ 12 axay 33 azz ( ) mOSxSLOSySLOSzS&t>0 Initial Condition: T(x, y, z, 0) = T... (3.5a) Boundary Conditions: -k33a%z+h,(T—T,,)=q0 atz=0,forOSx2+y2Sr2,0StStp (3.5b) h,(T-T,,,,)=k333T8z z=0,forr2Sx2+y2Sl,0StStp (3.50) h,(T—T,,.,)=k333Taz atz=0,forxSl,ySl,t>tp (35d) 56 8%xz0atx=0andx=l,fort>0 (3.5e) a%y=0aty=0andy=l,fort>0 (3.51) h,,(T-T,,)=k3337az atz=8, fort>0 (3.5g) where k,-,- refers to the ijth component of the thermal conductivity tensor, T... the ambient temperature, it, the convection heat transfer coefficient of the buoyancy-driven flow at the top surface, hi, the convection heat transfer coefficient for the stably stratified air layer at the bottom surface, r the radius of the circular heated region, and t, the heating time. During parameter estimation, the thermal diffusivity components 0:,- (a;,- = kij lpc) are estimated instead of the conductivity components kij. Figure 3.2 illustrates the model of the specimen geometry. The specimen is 40 mm by 40 mm by 3.25 mm thick. The heated region on the t0p face has a 6 mm radius and is illuminated for 1 s. The magnitudes of the top and bottom surface convection and heating boundary conditions are set as known, fixed parameters that are based on values obtained from an experimental apparatus used to measure the thermal diffusivity of slab specimens. Temperature sensors are located at four positions on the bottom face of the specimen: one sensor is at the center of the illuminated area, two of the sensors are located 9 mm from the center sensor along the coordinate axes defined by the plane of the specimen, and the final sensor forms the fourth comer of a square 9 mm to a side. The four sensors are named as shown in Figure 3.2. The finite difference model operates within a parameter estimation program that is based on a Levenberg-Marquardt iterative nonlinear least squares fitting routine (Press et al., 1992). This program also calculates show that an 81 x 81 x 41 node mesh produced grid independent temperature histories to within 0.001 K. 57 corner y—sensor SGHSOI’ D O Figure 3.2 Schema of the top surface of the specimen illustrating the illumination area. Top and bottom faces are subject to specified constant and uniform convection coefficients. Heating on the top face is of known power and duration; lateral edges are adiabatic. The temperature sensors are located on the bottom surface of the specimen at the comers of a 9 mm square that is aligned with the coordinate axes and has one sensor at the center of the bottom face; the projection of these sensor positions are shown on the top surface of the specimen. The principal axes of a, x’ and y’, are oriented at an angle B from the experimental axes; B is assumed unknown at the start of parameter estimation. Optimizing the duration of the experiment requires a criterion to judge the optimality of the design. A convenient optimization criterion for inverse heat transfer problems is to minimize the hypervolume of the confidence interval in the estimated parameters as measured in the covariance matrix (Beck and Arnold, 1977). Of the several methods available, D-optimum design is suited for estimation of multiple parameters because the 58 maximization of the D-optimum criterion, the determinant of the Fisher information matrix A, results in the minimization of the hypervolume of the confidence region about the estimated parameters (Beck and Arnold, 1977; Mejias et al., 1999; Emery, 2001 ). The Fisher information matrix is composed of scaled sensitivity coefficients F (F = p(8T/dp), where p is an estimated parameter and F has units of temperature) that indicate the relative change in the temperature field at a point with variation in a given estimated parameter. Comparing different experimental designs requires a non-dimensional definition of the Fisher information matrix, which may be written as <0 — 1 .irf[p an“, BEN—1 Yd! (36) '"'" M4, 3:] ,=0 map", "8p, Tm, ' where M is the number of sensors used (four here), If the total time over which data are collected, pm and p" the parameters being estimated, T, the temperature at sensor 3 as calculated by the model using the current estimate for p,, and Tm the highest temperature reached during the experiment. Note that the first two parenthetical terms inside the integral are scaled sensitivity coefficients. These terms are already calculated during the Levenberg-Marquardt parameter estimation, thus calculation of (I), and therefore A, requires little extra computational time. To investigate the effects of measurement time and random orientation on the ability to resolve components of a, temperature histories were calculated for a specimen possessing an out-of—plane diffusivity of 1.1 x 10'7 m2/s. The axial in-plane diffusivities, 0111 and 022, varied depending on the orientation of the principal axes of a, from a minimum of or" = 1.317 x 107 m2/s to a maximum of a. = 1.883 x 10'7 mz/s. The 59 corresponding shear diffusivity (002 = 021) varied from zero, where B = 0°, to a maximum of 0.282 x 10'7 mzls at B = 45°. The magnitudes of the components of a are on the order of those measured in manufactured PVC (Doss and Wright, 2000) or in loaded deformable elastomers (LeGall, 2001; Venerus er al., 2001). The values of on 1, (122, and 0112 are related to B and the principal in—plane diffusivities on and an by (1 +0. (I. ‘11 0L11 =(—I-§—Il]+[I—2£)'COS(B) (3.7) (1 +0. 0. —0. 0L22 = [I—ZJI-j—[iz—HJ'COSW (3.8) 0. —0. . a..=[I—21—I)-sm0) (3.9) Representative values of or; 1. 022, and 0112 are shown in Table 3.1. At the principal orientation, where B = 0°, the values of the in-plane axial diffusivities are at a maximum or minimum, respectively, and the shear diffusivity is zero. As B increases, the magnitudes of the axial diffusivities tend towards the median value of the principal in- plane diffusivities 011 and or" while the shearing diffusivity increases. When B = 45°, the shearing diffusivity is at a maximum and the in-plane axial diffusivities are equal to the median value. For 45° < B < 90°, the shearing diffusivity decreases while the axial diffusivities again approach maximum or minimum values until B = 90°, where 01 attains the same numerical values as initially, except that the values of or” and 0122 are reversed. Temperature histories were generated at a sampling rate of 120 Hz over 150 seconds for the five values of B shown in Table 3.1. Three thousand data points were used for each case, resulting in six combinations of sampling rate and measurement duration ranging 60 Table 3.1 Representative values of B, on 1. 022, and 012 used in the analytical study. Test B (1.11 x 107 (mzls) 02 x 107 (mZ/s) (112 x 107 (m2/s) 1 0° 1.883 1.317 0.000 2 15° 1.844 1.356 0.141 3 225° 1.800 1.400 0.200 4 30° 1.741 1.459 0.244 5 45° 1.600 1.600 0.282 from 25 s at 120 Hz to 150 s at 20 Hz. Noise was added to each temperature history in order to simulate the experimental measurements that would normally serve as input for thermal property estimation. The noise had a Gaussian distribution with zero mean and a standard deviation of o = 0.01 K, which is on the order of the noise measured using IR thermography. The resulting thirty sets of data served as input for parameter estimation where the four non-zero components of thermal diffusivity were estimated, sensitivity coefficients were calculated for each estimated parameter at each of the four sensor positions, and A was calculated to provide a relative measure of the amount of information obtained from each simulated experiment. Investigation of the error inherent in assuming in-plane orthotropy for samples having a randomly oriented a required smaller variations in 7 than used in the above tests, therefore a total of 12 data sets were calculated for 0° S B S 45°. These temperature histories were sampled at a rate of 30 Hz over a duration of 100 s for a total of 3000 data points per sensor. These tests were used to develop sensitivity coefficients for each sensor position as well as to serve as input for parameter estimation. Noise, with the same attributes as before, was again added to each temperature history. To enforce 61 orthotropy during parameter estimation, the value of the shearing diffusivity was fixed to zero during the estimation process, leaving only the three axial diffusivities to be estimated. Once again, A was calculated to serve as a metric of comparison for the amount of information obtained from each case. 62 3.2 Results For each orientation of or listed in Table 3.1, the error in the estimated diffusivity, e = (05- 0to)/010, where as is the estimated diffusivity and org the input diffusivity for the calculated temperature profiles, was less than 0.1% for the axial diffusivities and less than 1% for the shearing diffusivity. The fit between the simulated temperature histories and the inverse solutions generated using the estimated diffusivities was within the standard deviation of noise in the data. Figures 3.3a and 3.3b show representative temperature histories and the corresponding inverse solutions for case 1 (B = 0°, where the principal axes of a. and physical axes of the specimen coincide) and case 5 (B = 45°, where the shearing diffusivity reaches a maximum), respectively. The inverse solutions fit the simulated data well as evidenced by residuals (defined as the difference between the modeled and measured temperature histories) that were computed for each case. As expected, the residuals oscillate around zero with a magnitude on the order of the noise in the input data. The x- and y-temperature histories overlap each other since the in-plane axial diffusivities will be equal for the orientation corresponding to the maximum shearing diffusivity (Figure 3.3b). The sensitivity coefficients at each of the four sensor positions in case 1 (Figure 3.4) and case 5 (Figure 3.5) exhibit uncorrelated, nonzero behavior, with the exceptions of the an coefficient in case 1 and the or, 1 and 022 coefficients at the center sensor in case 5. Recalling the definition of the scaled sensitivity coefficient F, when the shearing diffusivity has zero magnitude, as it does for the B = 0° orientation (Figure 3.4), the sensitivity coefficient should be equal to zero over the entire duration of the experiment. Note how the 012 sensitivity coefficient has a zero 63 x-sensor center sensor y-sensor 0 I Icomer sensor _1 . , . , . , f , . 0 20 40 60 80 100 time (s) 5 , ('9) center sensor 4 - 3 —4 :x: 2 - ® X-SBDSOI’, y-sensor HAwAw comer sensor -1 ' 1 * 1 * 1 ' I ‘ O 20 40 60 80 1 00 time (s) Figure 3.3 Simulated temperature excess, 9 = T-T... and inverse solutions for (a) orthotropic test (case 1) and (b) maximum shear test (case 5). Note that for case 5, the x- and y-sensor temperature histories are identical as a result of 0m and 0122 being equal. 64 0.8 0.16 ...... 0.4 I '0'" ...~...“-..... "31:... ' ‘ g ‘ a" " Q 0.08 o.’ V 0.2 v ,9 l— l. c 0 0 0.04 ‘ c; -0 2 ~ 0 00 .J -0 4 . . r . -004 . . . . 0 20 40 60 80 100 333 0 20 40 60 80 100 time (s) ......... at 1 time (s) ------- a22 (c) ....... - .12 (d) 6 0.8 0.6 ‘ 0.4 " ........... g ........ ,_ 0.2 0.0 emu“. -o.2 . “New." -2 . . . . .04 . . . . 0 20 40 60 80 100 0 20 40 60 80 100 time (s) time (s) Figure 3.4 Scaled sensitivity coefficients for case 1: (a) y- sensor, (b) corner sensor, (c) center sensor, ((1) x-sensor. The solid line represents F for 033, the dotted line F for or, 1. the dashed line F for 0122, and the dot-dash line F for (112. Time is not scaled, as the choice of appropriate scaling parameters is not immediately obvious. For example, if the specimen thickness and corresponding axial diffusivity (133 are used as scaling parameters, than 1 second would correspond to a Fourier number of 0.0104; however, if either of the principal in-plane diffusivities are used with the specimen width as scaling parameters, 1 second would correspond to a Fourier number of either 8.23 x 10'5 or 1.18 x 10 . 65 (a) (b) 0.8 0.16 0 6 I'l"..........“~. .. 012 1 '/.—”-~\\ Q4 . ." ........ / \3.‘ 2 '0 ‘. ..... A 0.08 .1 '/ ’.'.’ ............. Q: I: 0.2: 5 // /,.’°’ L— 0.04 ‘ r 0.0 r — . f\ -0.2 ,, 0.00 ‘ -0.4 . . . v .004 r . . . 0 20 40 60 80 100 a33 0 20 40 60 80 time (s) ......... a11 time (s) ------- a22 (C) ------- - 312 (d) 6 0.8 06 4 .................... 4 . ........... 0.4 ............ Q q g ..... ,_ 2 L- 0.2 o 0.0 fl... \.._ 2.- 0.2 “‘~ ....... ,,__ __ -2 . 4 . . . .04 . . . . 0 20 40 60 80 100 0 20 40 60 80 time (8) time (s) Figure 3.5 Scaled sensitivity coefficients for case 5: (a) y-sensor, (b) comer sensor, (0) center sensor, ((1) x-sensor. The solid line represents F for 0133, the dotted line F for on 1. the dashed line F for 022, and the dot-dash line F for 002. value at all sensor locations in case 1 (Figure 3.4), while there is a change in the magnitude of the m2 sensitivity coefficient at each sensor position in case 5, where the shearing diffusivity is a maximum (Figure 3.5). The amplitude of the 012 sensitivity coefficient increases at each sensor location as B and the shearing diffusivity increase. In order to estimate a particular parameter to a high degree of confidence, the magnitude of the corresponding scaled sensitivity coefficient should be on the order of the temperature 66 100 100 signal at that point and the behavior of all of the sensitivity coefficients at that location should be uncorrelated. Over time, large variations in coefficient magnitude are preferable to relatively constant trends. The magnitude of the largest sensitivity coefficient at each sensor position is on the order of the temperature signal at that sensor. The sensitivity coefficients reach their maximum values at different times, which is preferable when estimating independent parameters. Assuming thermal orthotropy (i.e., assuming B = 0°, leading to shearing terms of zero magnitude) for a specimen having principal values of a = [1.10, 1.88, 1.32] x 10'7 m2/s with randomly oriented in-plane axes results in small errors in the estimated diffusivity values over the range of in-plane orientation 0° S B S 45° (Table 3.2). Note that the error in the estimated diffusivity value never exceeds 0.4%. The small error in the axial diffusivities results from the shearing diffusivity value being at most 18% of the value of the mean axial diffusivity. This error increases as the magnitude of the shearing diffusivity increases, e.g., this error increases by an order of magnitude when the maximum shearing diffusivity is doubled as shown in the first half of Table 3.3. Increasing the maximum shearing diffusivity to 72% of the median axial diffusivity value increases the maximum estimated axial diffusivity error to greater than 9% (Table 3.3). Also included in Table 3.2 is the normalized x2 for each test, which gives an indication of the quality of fit between the input and inverse model temperature histories. The x2 has units of temperature squared and is defined here as N 2 1 X2(Pi)=-A72[Tj “Yj(Pi)] (3.10) i=1 67 where N is the number of data points, p,- the parameters being estimated, T,- the measured (simulated) temperatures, and y,-(p,-) the model value corresponding to the jth temperature. Based on this definition, x2 should have a value on the order of the square of the magnitude of the noise in the measurements. Assuming that the estimated parameters are known precisely and that the noise has a Gaussian distribution with a zero mean and a magnitude on the order of $0.01 K, the value of x2 should be 4 x 10“1 K2. As B increases for a given specimen, the shearing diffusivity increases and the simulated material behavior becomes less orthotropic. While the estimates of the axial diffusivities change little with increasing B, the overall )8 increases, meaning that the quality of the fit between the temperature histories of the model and the input data decreases. Note in Tables 3.2 and 3.3 that this error increases with increasing B, corresponding to increasing shearing diffusivity. This degradation of the overall ability of the model to fit the input data is manifested in the comer temperature sensor (Figure 3.6). Note how the model trace no longer lies within the data scatter, indicating that data obtained from the comer temperature sensor may be used to verify whether the temperature sensor placement on an anisotropic specimen is coincident with the principal directions of diffusivity, and thus, whether an orthotropic thermal response may be assumed during the experiment. These data suggest that when the normalized x2 approximately 20 times the level of the noise squared (x2 = 84.3 x 104 K2), estimates of the orthotropic components of thermal diffusivity may be in error by 1% or more. 68 Table 3.2 Error in estimating diffusivity when assuming orthotropy (i.e., neglecting 002) for an anisotropic material with principal axes oriented by B with respect to the measurement axes. The maximum shear diffusivity for this material is about 18% of the median axial diffusivity value. Case (3 e in 6133 (%) e in or” (%) e in (122(%) 22 x104 (K2) 6 0° 0 0 0.08 3.95 7 25° 0 0 0.15 4.24 8 75° 0 0 0.23 6.19 9 11° 0 0 0.22 8.65 10 15° 0 -0.05 0.22 12.5 11 18° 0 -01 1 0.22 16.0 12 225° 0 -0.17 0.14 21.7 13 26° 0 -0.17 0.07 26.4 14 30° -0.09 -0.23 -007 31.9 15 35° -0.09 -035 .007 37.0 16 40° -0.09 -03 -019 41.1 17 45° -0.09 -0.38 -025 42.1 Table 3.3 Error in estimating diffusivity when assuming orthotropy for an anisotropic material with principal axes oriented by B with respect to the measurement axes. The maximum shear diffusivity for this material is about 35% of the median axial diffusivity value for tests 18-22 and about 64% of the median axial diffusivity value for tests 23-27. Case (3 e in (133 (%) e in or“ (%) e in 6122 (%) x2 x104 (K2) 18 0° 0.00 0.00 0.10 3.98 19 15° .0.09 -057 0.00 40.3 20 225° -027 -1.10 -075 84.3 21 30° -0.55 -1.65 -152 134. 22 45° -0.73 -231 -225 191. 23 0° 0.00 0.00 0.00 4.02 24 15° -0.64 -1.49 -307 127. 25 25° -l.64 -359 -7.52 416. 26 32° -2.45 -5.62 -9. 19 670. 27 45° -3.18 -8.88 -8.81 924. 69 Optimizing the duration of the experiment for a given (1 having principal values of [1.10, 1.88, 1.32] x 10'7 mzls and a heating time of l 8 indicates that the maximum A is obtained at 75 8, assuming that the amount of sampled data is held constant between tests. Figure 3.7 shows the variation of normalized A as a function of measurement duration. Note that A/Amam is maximized for an measurement duration of between 75 and 100 s. The only case with a distinct maximum in A/Amax is for B = 0°, where the maximum lies at 100 s. Non-normalized values of A led to a maximum at 75 s for the orientation angle [3 = 45°. 70 1.0 x-sensor, y-sensor -' «f ,J'xa.§£t...4.w.u.s {’19. 0.8 7 1 0.6 ‘ A it Q corner sensor 1 _ 0.2 7 0.0 ‘ r’" ‘ , 'O.2 ' l ' I r l ' 1 0 20 40 60 80 100 time (s) Figure 3.6 Input data and model tits at the x-, y-, and comer sensors for case 5 (B = 45°) where orthotropic behavior is assumed during parameter estimation. The solid lines represent the model fits for the simulated temperature history at the indicated sensor. The poor fit between the data and the model temperature histories at the comer sensor is due to the assumption of no shearing diffusivity. The temperature responses at the x- and y- sensor locations overlap since the x- and y—diffusivities are equal. 71 1.0 7 ' 9 U 0.8 ‘ l 0.6 7 <8 i 0 a 0.4 7 5 O O 0.2 " Q? 0.0 “ ' O T l ' I ' I T l 7 l r 0 25 50 75 100 125 150 time [s] Figure 3.7 Normalized A as a function of measurement duration for cases 1 through 5. Empty circles represent data taken for B = 0°, triangles B = 15°, squares B = 22.5°, diamonds B = 30°, and filled circles B = 45°. 72 3.3 Discussion For the cases where all four non-zero components of a were estimated, the maximum error in the estimated values of the axial components of a was within 0.1% of the input value, while the maximum error in the estimated shear diffusivity was within 1% of the input value. The error in the estimated axial diffusivities is on the order of that seen by Graham et al. (1999a), where the estimated parameters were the principal diffusivities and the orientation angle B. In that study, the maximum error in the estimated diffusivities was 0.2% of the input value. Here, the shear diffusivity error increased from 0.1% to nearly 1% for very small B, where the value of the shearing diffusivity approached zero, magnifying the effects of small estimation errors. Inspection of F at each sensor location for the four estimated components of (1 showed uncorrelated behavior for each non-zero component of (1, provided that data from all four locations are taken into account. Neglecting the behavior of the om and an sensitivity coefficients at the center sensor, each sensitivity coefficient reaches its maximum at a different sensor position, and it is at that position where the most information is obtained concerning the parameter associated with each coefficient. For example, one expects that much of the information about 0133 is gained from the center sensor during the first 20 to 30 seconds of the experiment (Figure 3.4). Measurements on PVC conducted via a three-sensor format have shown this to be the case. The behavior of the sensitivity coefficients shows that the x- and y-sensor locations provide most of the information used to determine an and (122. Because the or“ and org; sensitivity coefficients have nearly the same behavior at the center sensor position (or even identical behavior when their respective diffusivities are 73 equal, as shown in Figure 3.5a), it would be difficult to determine each parameter independently using that sensor position alone, even though the magnitudes of these coefficients are higher at the center position than any of the other sensors. The responses of these sensitivity coefficients are clearly independent, however, at either the x- or y- sensor locations (Figures 3.5c and 3.5d). This result indicates the importance of using the appropriate sensor locations to estimate multiple parameters in thermophysical property measurements. The comer sensor location delivers the most information about 0112, because the (112 sensitivity coefficient has its highest value there. Additionally, the magnitude of the 0112 sensitivity coefficient is on the order of the other sensitivity coefficients at that location only. Such behavior indicates that the four-sensor array described in this study provides adequate information for resolving the four components of 0. that are of interest here. Furthermore, the location of the sensors affects the ability to estimate the components of a. If the two in-plane axial diffusivities are to be the only estimated parameters, and the center (Figure 3.5a) and corner (Figure 3.5b) sensor locations were used to estimate these parameters, the correlated behavior of the 0m and an sensitivity coefficients at both locations would make the independent estimation of each parameter difficult, or even impossible. Because specimen orientation may be unknown prior to the experiment, sensor placement on orthogonal axes may ensure the best chance of obtaining uncorrelated sensitivity coefficients for the parameters of interest. Estimating uwhen orientation along principal axes has been assumed for a specimen that is randomly oriented resulted in errors in the estimated axial diffusivities below 1% 74 irrespective of the specimen orientation when the maximum shearing diffusivity of the specimen was about 18% of the median axial diffusivity. This error increases as the magnitude of the shearing diffusivity increases, as shown in Table 3.3, however, the magnitude of anisotropy typically seen in experiments on elastomers is closer to that of the low-shearin g diffusivity cases listed in Table 3.2. The in-plane to out-of-plane anisotropy ratio calculated in residually stressed PVC was 1.25 (Doss and Wright, 2000), while recent work on loaded elastomers reported anisotropy ratios below 1.2 for stretches of up to twice the original specimen length (Venerus et al., 2001). Here, the ratio between principal in-plane diffusivities was 1.43 for the low shear cases (Table 3.2). The anisotropy ratio will increase as the maximum shearing diffusivity increases, provided that the median axial diffusivity is held constant (Table 3.3). Thus, while a high level of anisotropy may affect the accuracy of results obtained under assumptions of orthotropy, such levels are unlikely to be encountered except in composites that are manufactured to have such properties, therefore, orthotropy may be assumed during property estimation for materials possessing low levels of anisotropy with little loss of accuracy. Still, if high levels of accuracy are needed during property estimation, it may be advantageous to measure 0. at orientations that are not aligned with the principal axes of or once a specimen is determined to exhibit in-plane anisotropy. Figure 3.8 shows the variation of Awith B for a given a. Note that A increases as B increases, despite the number of data points being the same for each B. While the error in the estimated diffusivities is always less than 1% for the lowest shearing diffusivity cases (Table 3.2), suggesting that specimen orientation had little effect on parameter estimation, the amount of information obtained during each test as quantified by A increased by at least an order of magnitude 75 when the specimen orientation was rotated away from the principal thermal axes. Because maximizing A minimizes the hypervolume of the confidence region of the estimated parameters (see Mejias et al., 1999), A is inversely related to the uncertainty in the estimated parameters, therefore, parameter estimation experiments performed on anisotropic specimens oriented at B = 45° with respect to the principal axes of a should yield more precise results, if the orientation is known prior to testing. Given that the boundary conditions here were assumed perfectly known, the effects of additional information obtained at the off-axis orientations may help increase accuracy in actual experiments where only approximations of boundary conditions may be known or significant uncertainties exist in geometric parameters such as sensor location or specimen thickness. Investigation of the optimal measurement duration for a particular or indicated that the optimal duration was between 75 s and 100 s for each orientation angle B that was studied. A prior study on orthotropic materials possessing identical in-plane axial diffusivities showed that the optimal measurement duration occurred when the in-plane temperature histories reached their maximum values (Davis and Wright, 2005). Other prior studies have argued on physical grounds that the optimal measurement duration coincides with the maximum in-plane temperature rise. Doss (1998), for example, found that the estimated diffusivities om of A181 304 stainless steel varied by more than 5% when the experiment duration was nearly half the optimal value, but the estimated values were within 2% 0f the reference value when the optimal measurement duration was used. l/eGall (2001) found that the estimated properties of plate glass varied by less than 0.5% 76 when the optimal measurement duration was used, but when the measurement duration was altered by i50%, the scatter in the estimated diffusivities increased to 5%. The x- and y-sensor temperature histories in this particular case reached their maximum values between 75 s and 100 s for the B = 0° orientation. Since the alignment of the specimen axes with the principal axes of (1 leads to orthotropic behavior at this orientation, these results resemble those from prior studies, but with the extension that the orientation angle has little effect on the optimal measurement duration. Based on the results of the prior study, one might expect the optimal measurement duration to be the same as the longest sensor rise time, which would occur in this study at the comer sensor, but actually the optimal measurement duration occurs at an earlier time. The interaction of the components of the diffusivity tensor with the fixed boundary conditions renders the problem complicated enough that computational simulation is required to find the optimal experimental time. 77 10'3 i 10'4 ‘3 o 10'5 10'6 10'7 10'8 10'9 “a 1 0‘1 0 1 ' l . 1 ' 1 ' I ' O 10 20 30 4O 50 Figure 3.8 Variation of A for an or having principal values of [1.10, 1.88, 1.32] x 10'7 m /s. The amount of information obtained from a measurement increases as the specimen orientation shifts away from the principal axes of a, Note that for B = 0° and 25°, A is six and two orders of magnitude smaller, respectively, than A at B = 45°. 78 3.4 Closure An analytical study was performed where the four non-zero components of the diffusivity tensor (1 were estimated for a material possessing random in-plane anisotropy on the order of certain manufactured or mechanically loaded elastomers. Sensitivity coefficients were calculated for each of the non-zero components of the a tensor at four sensor locations. The amplitude of the sensitivity coefficients suggest that the signal received at each sensor is sufficient for the estimation of the parameters, while the character of the coefficients over time suggests that the parameters are independent. Small values of in-plane anisotropy lead to negligible (< 0.4%) errors in the estimated parameters when the shearing diffusivity is neglected. Larger differences in anisotropy require the estimation of the shearing diffusivity, which is possible with the addition of a fourth sensor to the extended flash diffusivity method (Doss and Wright, 2000). This sensor may be used to verify assumed thermally orthotropic behavior in materials. 79 APPENDICES 80 APPENDIX A NOTES CONCERNING IR THERMOGRAPHY 81 The use of IR thermography as a temperature measurement technique for extended flash thermal diffusivity measurements is particularly desirable due to the flexibility it provides the experimenter for positioning of the temperature sensors. Additionally, thermo-optical methods of temperature measurement make no contact with the specimen surface that could disturb the resultant temperature field, thus eliminating a source of error in the experiment. However, the use of IR thermography presents its own challenges that must be considered if such a method is to be used in the lab. Presented here are suggestions concerning (a) the general use of IR thermography for temperature measurement and (b) specifically, the use of our own IR camera in the data acquisition and subsequent parameter estimation procedure. The camera used for making experimental temperature measurements in these studies is an Indigo Merlin Mid-IR (3 — 5 pm) model that is Stirling-cycle cooled. The camera uses a 30 um x 30 um indium antimonide (InSb) detector to record a 320 x 256-pixel field at a frame rate of up to 60 Hz. The camera lens currently in use has a 50 mm focal length, a field of view of 11° x 8°, and an instantaneous field of view of 0.6 milliradians. The response time of the detector, as measured in cameras of similar wavelengths, is on the order of several microseconds (Torres 42! al., 1990). The camera comes factory calibrated to measure temperature over the range of 0-350 °C, but can be adjusted via software to handle temperatures up to 2000 °C if necessary. Typically, the temperature range of interest for our measurements is 20-40 °C. The noise seen in spot camera measurements is approximately $0.01 K, about an order of magnitude less than that seen when using E- type thermocouples with commercial calibration. The camera pixel size depends on the 82 Optics as well as the distance between the specimen and lens and is 210 um square in the current test configuration. The camera captures full-field data at a constant frame rate of 60 Hz. Using the accompanying software, it is possible to capture data at slower rates by saving every 11th frame, leading to successive capture rates of 30 Hz, 15 Hz, 12 Hz, 10 Hz, etc. Slower capture rates are preferable for lower diffusivity materials such as PVC, polyurethane rubber, or plate glass. The maximum rate of 60 Hz is sufficient for capturing data from a 1.5 mm thick AISI 304 stainless steel plate, but 3-D flash tests on such a specimen yield only about 350 data points if the data is sampled until the off-axis temperature sensors (assumed to lie ~8-10 mm from the center sensor) reach their maximum temperature (~6 s @ 60 Hz). The center-sensor temperature rise occurs in 2-3 s for materials with diffusivities on the order of steel; should the desire to examine a material with a higher diffusivity arise, a camera with a faster frame rate should be used for data capture. As noted in this dissertation, there is a random offset between the laser and camera triggers due to the fact that the camera does not synchronize precisely with the laser trigger. For elastomers, this offset is negligible, causing little to no change in any estimated parameters, but for higher diffusivity materials like steel, several tests must be run and subsequently averaged to arrive at an appropriate value for the out-of—plane diffusivity. As for sensor positioning, the act of moving the off-axis sensors further away from the center point results in a slower rise time, allowing more data points over the temperature history, but also a lower maximum signal and a poorer signal to noise ratio. It is possible that the constant convection boundary condition modeled on the top surface (and estimated as a nuisance parameter) breaks down near the edges where a convective plume would likely pull cool air in from the surroundings. As a rule of thumb, the outer 83 5-7 mm of the specimen probably shouldn’t be used for data sampling with the finite difference models included in this dissertation. Additionally, the specimen holder itself may cause parasitic heat losses. Originally, one of the sensor positions used in the study was located between the illuminated area and the specimen holder (which clamps to the specimen at one point via a screw). The metal holder was found to affect the temperature response at this sensor location, causing a lower temperature signal than what was expected. The responses at the other off-axis locations yielded an isotropic response. It is recommended that sensor positions always be located away from the point where rigid specimens are clamped to the device. The detector records the infrared luminosity of the target and converts that number to a temperature value. Unfortunately, the target luminosity can be biased by reflected infrared light from elsewhere. The entire testing apparatus has been covered with a tarp that is secured to a rigid frame to minimize intrusion of light from outside sources, but light still leaks in from around power cords, and the camera itself generates heat which can unevenly warm the apparatus and surroundings. Surface reflectivity effects were seen in the full field images when testing barren aluminum and steel plates, and minor effects were also seen when testing PVC plates. To counteract the reflectivity, flat black stove spray paint was used to coat the top (illuminated) and bottom (interrogated) surfaces of the specimens. This not only reduced the surface reflectivity of the specimen, but also the transmissivity of the laser light through translucent specimens such as glass. The thickness of the paint is on the order of microns, several orders of magnitude smaller than the specimen thickness, and thus has a negligible effect on the estimated parameters 84 of the experiment. In prior experiments, a similar paint was used to eliminate transmission of laser light through a glass specimen while using a fixed thermocouple probe for temperature measurement. At the time, it was believed that the laser energy was completely absorbed at the coating surface, but IR camera tests on the plate glass specimen showed that the laser light penetrates through both painted surfaces resulting in a small but noticeable signature in the center sensor temperature history that ceases when the laser pulse ends. Thus, the paint is opaque to visible light, but lets infrared light through. Multiple coats of the paint do not remedy the problem. A similar signature is not seen in the center sensor temperature history of any metal or elastomeric specimens. Should the desire to test a translucent or transparent material arise, another material, preferably one that may be applied via spraying so as to minimize the amount of deposited material, must be used. Since the stove paint appears opaque but actually transmits a fraction of the incident infrared light, possible coatings for the specimen should be tested using the camera to check for a signature from the laser. The laser should be tuned to its highest power for a 1-2 s length pulse time, depending on the thickness and estimated properties of the material. If there is transmission, a single sawtooth-shaped spike will appear in the center temperature sensor temperature history at the end of the flash time. The spike seen during glass measurements was roughly one to two orders of magnitude greater than the ambient noise. If transmission is absent, the temperature rise at the center sensor should be smooth (within the bounds of the experimental noise). 85 When data is collected, it is saved in a Dynamite movie file. This file can then be saved frame by frame. These single frames may be opened in the Thermogram software, and it is from here that temperature histories at specific locations may be pulled using the Plot tools. After the appropriate sensor areas are selected, temperature histories may be pulled from each frame and displayed on a visual plot. This data may then be saved as an ASCII file, and this file is used as input for the parameter estimation program. Currently, the Marquardt program reads in 3-sensor ASCII data. The format of the data line from the ASCII file is “date, time, sensorl, sensor2, sensor3”, however, the file formatting that the camera uses depends on the time at which the camera is being used -— different times of day lead to different spacings between the time and the temperature values, and thus when running news tests on a specimen, the data files that are read in must be checked to ensure that the correct formatting is used for the input statements. The parameter estimation program pauses after reading in the data file, giving the experimenter a chance to check the input data, which is echoed to the “dataout” file. If the data file is not read properly, negative temperatures will appear, or the temperature response at a given sensor location will suddenly jump up and down by tens of degrees. To correct this, the FORMAT statement that is used to read in the data inside the “readit” subroutine in the Marquardt program must be altered. Comment lines inside the program list the changes necessary to correct for several common formatting errors. To approximate point-wise temperature measurement, the Thermogram software is used to record data at three points using the “rectangle” tool to record average temperatures over a 3 X 3-pixel area. This sensor area is of sufficient size that the detector will capture 86 the true temperature of the position of interest yet minimize noise. This 3 x 3-pixel array produces a sensor area of about 0.6 mm by 0.6 mm. While this area is smaller than the surface area of the E-type thermocouples previously used for the fixed-position temperature probe, there is still some uncertainty in the position of the sensor. Sensor positions should be determined using a reference material with known thermal properties. Setting the thermal diffusivities as known quantities, estimate the off-axis sensor positions, incident heat flux and top surface convection coefficient using the Marquardt program. Once the sensor positions are confirmed, materials of unknown thermal properties may be tested. The clamp inside the testing device is set such that the bottom surface of the (rigid) specimen is always the same distance from the camera lens; as long as the camera and clamp are not moved, the pixel dimensions will remain the same and sensor positions do not need to be recalculated. 87 APPENDIX B ALTERNATING DIRECTION IMPLICIT (ADI) FINITE DIFFERENCE PROGRAM FOR MODELING ORTHOTROPIC HEAT CONDUCTION IN A SLAB SPECIMEN 88 The following pages contain FORTRAN code for an alternating direction implicit (ADI) finite difference program that models the diffusion equation and assumes thermal orthotropy in a slab specimen. The thermal diffusivity tensor (1 that allows an orthotropic response may be expressed as 111] 0 O _ 9= 0 922 0 (8.1) _ O 0 (133a The simulated specimen boundary conditions consist of: insulated edges, convection on the top and bottom surfaces (of different magnitude), and a top surface heat pulse approximately 5.3 mm by 5.3 mm with constant magnitude and finite duration. The estimated parameters include the three axial components of a, a scaled heat flux coefficient, and a scaled convection coefficient for the top surface. The bottom surface convection coefficient is assumed known. Most of the experimental and geometric parameters are adjusted via the setup subroutine that precedes the actual model routine. The output consists of an array containing the temperature history of the simulated specimen at three points on the bottom surface. 89 O ()(5 O ()(5 O ()(5 0 (3(3 0 C) O Implicit ADI model for IR camera data reduction. Model is used with Marquardt program to solve IHCP for a specimen of unknown thermophysical properties. This monolayer model has a rectangular, uniform surface heat flux on the top surface of specimen. All side boundaries are insulated. The top and bottom surfaces have convection BCs implemented due to unstably stratified air heating on the top and stabily stratified heating on the bottom (Htop > Hbottom). The bottom convection coefficient is assumed known. ******************* this routine sets up fixed parameters for the problem to be solved. I wanted to set all the variable in one program as opposed to looking all over creation to change the problem of interest. subroutine setup(IMAX,STOPITER,XAPT,YAPT) CHANGE model size here NMAX should equal NX times NY times NZ for this prog. useful combinations: (21,21 ,21,9261) (41,41,41,68921) (61 ,61 ,41 , 152561 ) integer NMAX,NX,NY,NZ parameter (NX = 41) parameter (NY = 41) parameter (NZ = 41) parameter (NMAX = 68921) real(8) zlength,xlength,ylength,xratio,yratio,ft real(8) dz,dx,dy,QIB,QJB,fliz,ALPHAZ,ALPHAX,ALPHAY,QEST,DHZ real(8) pcspec,flashtime,freq,dt,BOTCONV real(8) endtime,timecalc,STOPITER,scalz,scalx,scaly,SURFABS real(8) tdiff(7),tdiffo(7),dx2i,dy2i,d22i real(8) XAPT,YAPT,lxtc,lytc,xcorr,ycorr,Tmax integer QI,QJ ,nreada,dataswitch,PN X,PNY,PN Z,NTOTAL,N POINTS integer FLASH,IMAX,x(5),y(5),z(5),nread,XTC1,YTC1,XTC2,YTC2 integer noiseswitch,sensors character*12 fname common blocks common /GUESS l/ tdiff,tdiffo common IOTPT/ x,y,z,XTC1,XTC2,YTC1,YTC2,xcorr,ycorr common lPOSlT/ PNX,PNY,PNZ common /RFU QI,QJ ,QIB,QJ B 90 0000000 0000000 0 common IREADITT/ nreada,dataswitch,fname,scalz,scalx,scaly,fhz common IGEOM/ dx,dy,dz common IGEOM2/ dx2i,dy2i,d22i,dt common /ETC/ NPOINTS,NTOTAL,FLASH,endtime,pcspec common ITIME/ timecalc common IFOO/ nread common lFLASHIT/ flashtime common /ZL/ zlength common IOUT2/ noiseswitch common IFISI-I/ sensors,Tmax *model specimen property parameters (initial guesses)* 3-D thermal diffusivity, heat flux, convection ALPHAZ = 1.25D-07 ALPHAX = 1.25D-07 ALPHAY = 1.25D-07 QEST = 0.031 QEST = 0.02 DHZ = 1.25D-05 ALPHAZ = 4.1D-06 ALPHAX = 4.1D-06 ALPHAY = 4.1D-06 QEST = 0.01 DHZ = 1.0D-05 DHZ = 0.0 assume that pc of the specimen can be found thru DSC. glass pcspec = 2.268D+06 pvc pcspec = 1.540D+06 steel pcspec = 3.866D+06 rear face convection -- assumed known to reset to insulated boundary, zero out BOTCONV BOTCONV = 5.0 BOTCONV = 0.0 *model run parameters* sampling frequency, # of data points, overall calc. time fhz=15. fl'lZ = 30. 91 fhz = 60. nreada = 1190 timecalc = 18.01 flash time flashtime = 1.0 flashtime = 10.0 flashtime = 1.31 flashtime = 2.2 dataswitch: 1 = IR camera (3-spot) data, 2 = TC data, 3 = contrived test data 4 = forward problem dataswitch = 1 noiseswitch: 0 = no noise, 1 = add noise this should be set to "l" for the forward problem only! noiseswitch = 0 input (data) file name. fname = "st04c.asc" *location of 'output' nodes (in mm)* lxtc = 20.0 lytc = lxtc lxtc = 17.26 lytc = 16.23 *nodal geometry parameters* input dimensions in millimeters the model is nondimensionalized by the specimen thickness zlength = 3.355 zlength = 2.25 zlength = 1.50 xlength = 20. ylength = 20. zlength = l. xlength = 1. ylength = 1. sensors = 3 scaling parameters for TC data scalz = 0.00 scalx = 0.00 scaly = 0.00 surface absorption only for this model (for now) SURFABS=1.00 92 0 iteration parameters (max. # of iterations; convergence criterion for alphaz) IMAX=50 STOPITER= 1 .0D- 10 nondimensionalization xratio = xlength/zlength yratio = ylength/zlength dz = 1./dfloat(NZ- 1) dx = xratio/dfloat(NX- 1) dy = yratio/dfloat(NY- 1) dx21 = l./(dx*dx) dy21= l./(dy*dy) dzZi = 1./(dz*dz) write(*,*) xratio, yratio, dz, dx, dy heat flux/aperture parameters. The QI/QJ terms specify the highest node to which the unit heat flux will be applied. XAPT = 5.85 XAPT = 5.57 YAPT = XAPT QI = nint(XAPT/(dx*zlength)) Q] = nint(YAP’F/(dy*zlength)) QI = NX QJ = NY these terms add a portion of the total flux to the boundary of the heated surface -- needed if the actual geometry doesn't line up with the descretized geometry. QIB = (XAPT-(float(QI)*dx*zlength)+(0.5*dx*zlength))/(dx*zlength) QJB = (YAPT-(float(QJ)*dy*zlength)+(0.5*dy*zlength))/(dy*zlength) QIB = 0. QIB = O. if((QI.gt.NX).or.(QJ.gt.NY)) pause 'flux defined area too big' Output nodes are, by definition, usually on the bottom face of the specimen. Currently, (1) = top face center, (2) = bottom face center, (3) = xTC, (4) = yTC, (5) = center bottom outside edge of specimen. points x(3) and y(4) will be written over for the inter-nodal back face points x(1)=l x(2)=1 93 CO x(3) = O x(4) = 1 x(5) = NX y(1)= 1 W) = 1 y(3) = l y(4) = 0 W) = 1 2(1) = 1 2(2) = NZ z(3) = NZ 2(4) = NZ 2(5) = NZ PNX = NX PNY = NY PNZ = NZ NTOTAL = PNX*PNY*PNZ if (NT OTAL .ne. NMAX) pause 'NTOTAL does not equal NMAX' if (int(timecalc*fhz).gt.nreada) pause 'check nreada -- too small' freq = fhz*flashtime NPOINTS = int(timecalc*fl12) dt = ( l .lfreq) endtime = dfloat(dt*NPOINTS)*flashtime FLASH = int(freq) ft = float(FLASH)/fhz this section sets the estimated parameters into the array that will pass back and forth between the model and the MO reduction scheme. tdiffl I )=ALPHAZ tdiff(2)=ALPHAx tdiff(3)=ALPHAY tdiff(4)=QEST tdiff(5)=DHZ tdi ff(6)=SURFABS tdiff(7)=B OTCONV 94 this section takes care of inter-nodal IR camera query points lxtc = lxtc/zlength lytc = lytc/zlength XTCl = int(lxtc/dx) + l YTCl = int(lytc/dy) + 1 XTC2 = XTCl + 1 YTC2 = YTCl + 1 xcorr = l. - ((lxtc - (XTC1-1)*dx)/dx) ycorr = 1. — ((lytc - (YTC1-1)*dy)/dy) Tmax = 0. return end C****************************************************************** C this is the body of the bilayer model program subroutine model to alter the size of the model, update the parameters NX, NY, NZ and NMAX. To facilitate this, search through the program for the word "CHANGE" (as in model size). CHANGE model size here NMAX should equal NX times NY times NZ for this prog. integer NMAX,NX,NY,NZ parameter (NX = 41) parameter (NY = 41) parameter (NZ = 41) parameter (NMAX = 68921) *variable declarations* integer N TOTAL integer STEP,COUNT,FLASH,NPOINTS integer position,i,j,k,noiseswitch,idum real(8) flashtime,fluxswitch,endtime,kspec real(8) dt,dx,dy,dz,dx2i,dy2i,dz2i,time real(8) Fo,G 1 ,G2,Biot,q,Fol ,F02,Fo3,Bind,qnd,nBind real(8) AX,AY,AZ,BX,BY,BZ,CX,CY,CZ real(8) zlength,pcspec,backbiot,bBind,anind 95 O real(8) T(NX,NY,NZ),Tout(50000,5),noise real(8) al(NMAX),b1(NMAX),cl(NMAX),r(NMAX) real(8) a2(NMAX),b2(NMAX),c2(NMAX),u(NMAX) real(8) a3(NMAX),b3(NMAX),c3(NMAX) real(8) tone(NX,NY,NZ),ttwo(NX,NY,NZ) real(8) RA(9),RB(9),RC(9),RD(9),RE(9),RF(9),RG(9) real(8) RH(9),RI(9),RJ(9),RK(9),RL(9),RM(9),RN(9) real(8) RO(9),RP(9),RQ(9),RR(9) real(8) atry(7),tmod(50000) *** common blocks *** NOTE: temperature arrays are scaled, dimensionless temperatures. Do not treat T as T(deg. C)! NOTE: lengths are scaled too. Time is scaled! common blocks common [TEMPERATURE/ T,tone,ttwo,Tout common /PROPS/ Fo,G1,G2,Biot,q,qnd,Bind,nBind common IRVECT/ RA,RB,RC,RD,RE,RF,RG,RH,RI,RJ,RK,RL,RM,RN,RO common IRVECT2/ RP,RQ,RR,bBind,anind common IFLASHIT/ flashtime common /GEOM/ dx,dy,dz common /GEOM2/ dx2i,dy21,d22i,dt common /ETC/ NPOINTS,NTOTAL,FLASH,endtime,pcspec common lMAINPROG/ tmod,atry common /ZL/ zlength common /OUT2/ noiseswitch non-dimensional (specimen) parameter construction Fo = (atry(l)*flashtime)/((zlength/1000.)*(zlength/1000.)) G1 = atry(2)/atry(1) G2 = atry(3)/atry(l) G1 = 1. G2 = l. kspec = atry(l)*pcspec Biot = (atry(5)*(zlength/1000.)*pcspec)/kspec q = (atry(4)*pCSpec) backbiot = (atry(7)*(zlength/l000.))/kspec *** body of program *** from here to the "THIS IS THE LOOP" comment is only called once. 96 C C *** construct known coefficients *** initialize "building blocks" for specimen F01 = F0*Gl *dt*dx21 F62 = Fo*G2*dt*dy2i F63 = Fo*dt*d22i Bind = (2.*Fo*Biot*dt)/dz nBind = -l .*Bind bBind = (2.*Fo*backbiot*dt)/dz anind = -l.*bBind qnd = (4.*dt)/dz LHS coefficients internal nodes -- specimen AX = -1.*Fol AY = -l.*Fo2 AZ = -1.*Fo3 BX = 2.*Fol + 2. BY = 2.*F02 + 2. B2 = 2.*Fo3 + 2. CX=AX CY=AY CZ=AZ below are RHS coefficients first, specimen coefficients RB(1) = Fol RB(2) = Fol RB(3) = Fol RB(4) = 2.*Fol RB(5) = 2.*Fol RB(6) = Fol RB(7) = 2.*Fol RB(8) = Fol RB(9) = Fol RC(l) = 2.*F02 RC(2) = F02 RC(3) = F02 97 RC(4) = F02 RC(S) = F02 RC(6) = F02 RC(7) = 2.*F02 RC(8) = 2.*F02 RC(9) = F02 RD(l ) = 2.*Fo3 RD(2) = 2.*Fo3 RD(3) = F03 RD(4) = 2.*Fo3 RD(5) = F03 RD(6) = F03 RD(7) = F03 RD(8) = F03 RD(9) = F03 do i = 1,9 RA(i) = 2. - 2.*RB(i) - 2.*RC(i) - 2.*RD(i) RE(i) = 0. RF(i) = 0. RG(i) = 0. RH(i) = 0. RI(i) = 0. R] (i) = 0. RK(i) = 0. RL(i) = 0. RM(i) = nBind RN(i) = 0. RO(i) = 0. RP(i) = anind RQ(i) = 0. RR(i) = 0. enddo RE(2) = -2.*RB(2) RB(3) = -2.*RB(3) RB(5) = -2.*RC(5) RB(6) = -2.*RC(6) RE(8) = -2.*RD(8) RB(9) = -2.*RD(9) RF(2) = RB(2) RF(3) = RB(3) RG(5) = RC(S) RG(6) = RC(6) 98 0 RH(8) = RD(8) RH(9) = RD(9) RI(3) = -2.*RC(3) RI(6) = -2.*RD(6) 121(9) = -2.*RB(9) RK(3) = RC(3) RL(6) = RD(6) RJ (9) = RB(9) convection coeffs. RM(1)= 2.*nBind RM(2) = 2.*nBind RM(4) = 2.*nBind RN(8) = nBind RN(9) = nBind RO(6) = nBind RP(l) = 2.*anind RP(2) = 2.*anind RP(4) = 2.*anind RQ(8) = anind RQ(9) = anind RR(6) = anind construct coefficient vectors -- starting with internal nodes note that the RHS current node is constructed here as well for the internal current node ONLY. doi=1,NMAX a1(i)=AX b1(i)=BX cl(i)=CX a2(i) = AY b2(i) = BY c2(i) = CY a3(i) = AZ b3(i) = 82 99 c3(i) = CZ enddo construct A and C vectors on the boundaries top comers a1 (position(1,1,1 )) = 0. cl (position(],l, 1 )) = 2.*CX a1 (position(NX,], 1)) = 2.*AX cl(position(NX,1,1)) = 0. a1 (position(1,NY, 1)) = 0. cl(position(1,NY,1)) = 2.*CX a1 (position(NX,NY,1)) = 2.*AX c 1 (position(NX,N Y, 1 )) = 0. a2(position(1,l ,1)) = 0. c2(position(1,l,1)) = 2.*CY a2(position(NX,] ,1)) = 0. c2(position(NX,],1)) = 2.*CY a2(position(l,NY,1)) = 2.*AY 02(position(1 ,N Y, 1 )) = 0. a2(position(NX,NY,l)) = 2.*AY c2(position(NX,NY,1)) = 0. a3(position(] ,1 ,1)) = 0. c3(position(l,1,1)) = 2.*CZ a3(position(NX,1,1)) = 0. c3(position(N X, 1 , l )) = 2.*CZ a3(position(] ,NY,1)) = 0. c3(position(1,NY,l)) = 2.*CZ a3(position(NX,NY,l)) = 0. c3(position(NX,NY,l)) = 2.*CZ 100 c top and bottom edges doj=2,NY-l enddo al(position(1,j,1)) = 0. cl(position(1,j,1)) = 2.*CX a1(position(NX,j,l)) = 2.*AX c1(position(NX,j,1)) = 0. a3(position(1,j,1)) = 0. c3(position(1,j,1)) = 2.*CZ a3(position(NX,j,1)) = 0. c3(position(NX,j,1)) = 2.*CZ al (position( 1 ,j,NZ)) = 0. cl (position(] ,j,NZ)) = 2. *CX a1(position(NX,j,NZ)) = 2.*AX c1 (position(NX,j,NZ)) = 0. a3(position( 1 ,j,NZ)) = 2.*AZ c3(position( l ,j,NZ)) = 0. a3(position(NX,j,NZ)) = 2.*AZ c3(position(NX,j,NZ)) = 0. doi=2,NX-l a2(position(i,1,1)) = 0. c2(position(i,1,1)) = 2.*CY a2(position(i,NY,1)) = 2.*AY c2(position(i,NY,1)) = 0. a3(position(i,1,l)) = 0. c3(position(i,l,1)) = 2.*CZ a3(position(i,NY, l )) = 0. c3(position(i,NY,l)) = 2.*CZ a2(position(i, 1 ,NZ)) = 0. c2(position(i,l ,NZ)) = 2.*CY a2(position(i,NY,NZ)) = 2.*AY 02(position(i,NY,NZ)) = 0. 101 enddo a3(position(i, 1 ,NZ)) = 2.*AZ c3(position(i, 1 ,NZ)) = 0. a3(position(i,NY,NZ)) = 2.*AZ c3(position(i,NY,NZ)) = 0. top and bottom faces doi=2,NX-l enddo do j = 2, NY-l a3(position(i,j,1)) = 0. c3(position(i,j,1)) = 2.*CZ a3(position(i,j,NZ)) = 2.*AZ 03(position(i,j,NZ)) = 0. enddo side edges do k = 2, NZ-l a] (position( 1 ,1 ,k)) = 0. c1(position(l,1,k)) = 2.*CX a1 (position( 1 ,NY,k)) = 0. c1(position(l,NY,k)) = 2.*CX a1 (position(N X, l ,k)) = 2.*AX c1(position(NX,1,k)) = 0. a1 (position(NX,NY,k)) = 2.*AX cl (position(NX,NY,k)) = 0. a2(position(l,1,k)) = 0. c2(position(1,l,k)) = 2.*CY a2(position(l,NY,k)) = 2.*AY c2(position(1,NY,k)) = 0. a2(position(N X, l ,k)) = 0. 02(position(N X, 1 ,k)) = 2.*CY a2(position(NX,NY,k)) = 2.*AY c2(position(NX,NY,k)) = 0. 102 enddo side faces do j = 2, NY-l do k = 2, NZ] a1(position(1 ,j,k)) = 0. c1(position(1,j,k)) = 2.*CX a1(position(NX,j,k)) = 2.*AX cl(position(NX,j,k)) = 0. enddo enddo do i = 2, NX-l do k = 2, NZ-l a2(position(i,1,k)) = 0. c2(position(i, l ,k)) = 2.*CY a2(position(l,NY,k)) = 2.*AY 02(position(i,NY,k)) = 0. enddo enddo bottom comers a1(position(l,1 ,NZ)) = 0. c1(position(1,1,NZ)) = 2.*CX a1(position(NX,1,NZ)) = 2.*AX c1(position(NX,1,NZ)) = 0. al(position(1,NY,NZ)) = 0. cl (position( 1 ,NY,NZ)) = 2.*CX a1(position(NX,NY,NZ)) = 2.*AX c1(position(NX,NY,NZ)) = 0. a2(position(1,l ,NZ)) = 0. c2(position(1,l ,NZ)) = 2.*CY a2(position(NX,] ,NZ)) = 0. 02(position(NX,1,NZ)) = 2.*CY a2(position( 1 ,NY,NZ)) = 2.*AY 02(position(1,NY,NZ)) = 0. 103 00000 a2(position(NX,NY,NZ)) = 2.*AY 02(position(NX,NY,NZ)) = 0. a3(position(1,1,NZ)) = 2.*AZ c3(position(l,1,NZ)) = 0. a3(position(NX,1,NZ)) = 2.*AZ c3(position(NX, 1 ,NZ)) = 0. a3(position(1,NY,NZ)) = 2.*AZ c3(position(1,NY,NZ)) = 0. a3(position(NX,NY,NZ)) = 2.*AZ c3(position(NX,NY,NZ)) = 0. alter convection -- start it at t=0 (full face) do i = 1, NX do j = 1, NY b3(position(i,j,l )) = BZ + Bind b3(position(i,j,NZ)) = BZ + bBind enddo enddo reorder a,b,c vectors for the x and y directions so that they are in the proper order to be used by tridag. call reorder(a1,b1,c1,l ) call reorder(a2,b2,c2,2) initialize arrays and counters, switches T = 0. tone = 0. ttwo = O. COUNT = l fluxswitch = 1. THIS IS THE TIME LOOP. A, B, and C vectors have been constructed for all three directions. now we march through time implicitly, solving for the scaled temperature field at each time step so that we may compare the analytical solution with the experimental results taken from our data. 104 do while (COUNT.ne.0) if(COUNT.eq.(FLASH+1)) fluxswitch=0. STEP = 1 first time jump call rfill(r,STEP,fluxswitch) call flipr(r, 1) call tridag(al ,b l ,c 1 ,r,u,N TOTAL) call invpos(u,tone, 1) second time jump call rfill(r,STEP,fluxswitch) call flipr(r,2) call tridag(a2,b2,02,r,u,NTOTAL) call invpos(u,ttwo,2) third time jump-- this completes one full time step call rfi11(r,STEP,fluxswitch) call tridag(a3,b3,c3,r,u,NTOTAL) call invpos(u,T,3) output temperature at points of interest to necessary array call output(count) COUNT = COUNT + 1 if(COUNT.eq.(FLASH+1)) fluxswitch=0. 105 fourth time jump call rfill(r,STEP,fluxswitch) call flipr(r,2) call tridag(a2,b2,c2,r,u,NTOTAL) call invpos(u,tone,2) fifth time jump call rfill(r,STEP,fluxswitch) call tridag(a3,b3,c3,r,u,NTOTAL) call invpos(u,ttwo,3) sixth time jump-- this completes two full time steps call rfill(r,STEP,fluxswitch) call flipr(r, 1) call tridag(al,b1,cl,r,u,NTOTAL) call invpos(u,T, 1) outputtemperature at points of interest to necessary array call output(count) COUNT = COUNT + 1 if(COUNT.eq.(FLASH+1)) fluxswitch=0. seventh time jump call rfill(r,STEP,fluxswitch) call tridag(a3,b3,c3,r,u,NTOTAL) call invpos(u,tone,3) 106 c eighth time jump call rfill(r,STEP,fluxswitch) call flipr(r, 1) call tridag(a1,b1,cl ,r,u,NTOTAL) call invpos(u,ttwo, 1) c ninth time jump» this completes three full time steps call rfill(r,STEP,fluxswitch) call flipr(r,2) call tridag(a2,b2,c2,r,u,NTOTAL) call invpos(u,T,Z) c output temperature at points of interest to necessary array call output(count) ' COUNT = COUNT + l if((COUNT-l).ge.NPOINTS) COUNT = O c this enddo loops back for the next full time step enddo c output temperature profiles to a file c this section is used to get dimensional temperatures. doi=1,NPOINTS do j = 1, 5 Tout(i,j) = Tout(i,j)*((q*flashtime*1000.)/ + (pcspec*zlength)) enddo enddo C this will put the output in a useful form for the marquardt program. do i = 1, NPOINTS 107 00 0000000000 273 j = i+NPOINTS k = i+2*NPOINTS center TC tmod(i) = Tout(i,2) right now, set to old X TC tmod(i) = Tout(i,3) right now, set to old Y TC tmod(k) = Tout(i,4) enddo open file for model output (scaled) open (unit=92, file="Tmode1.txt", status = "unknown") noise = 0.01 if (noiseswitch.eq. 1) then call SYSTEM_CLOCK(idum) if (idum.eq.0) pause 'idum = 0; restart' idum = 4 write(*,*) 'idum = ‘,idum do i = 1,NPOINTS do j = 1,5 Tout(i,j) = Tout(i,j) + noise*gasdev(idum) enddo enddo confinue else pause 'error with noiseswitch' endif do count = l, NPOINTS time = COUNT*dt*flashtime write(92,273) time,Tout(count, 1 ),Tout(count,2),Tout(count,3), + Tout(count,4),Tout(count,5) enddo format(1X, F7.3, 2X, 5F8.3) close(92) return end 108 000 ************************************ **********Sprr0gramS*************** ************************************ integer function position(i,j,k) computes the position number of a 3-space node in a single vector array. Needed for compact matrix manipulation in tridag. common /POSIT/ PNX,PNY,PNZ integer i,j,k,PNX,PNY,PNZ position = ((1-1 )*(PNY*PNZ))+((j- 1 )*PNZ)+k return end function ************** subroutine reorder(a,b,c,dir) this subroutine reorders the a, b, and c vectors for the tridag routine to work properly. CHANGE model size here integer N MAX,N X,N Y,N Z parameter (NX = 41) parameter (NY = 41) parameter (NZ = 41) parameter (NMAX = 68921) integer i,j,k,m,dir real(8) a(NMAX),b(NMAX),c(NMAX) real(8) tempa(NX,NY,NZ),tempb(NX,NY,NZ),tempc(NX,NY,NZ) m = 1 call invpos(a,tempa,3) call invpos(b,tempb,3) call invpos(c,tempc,3) if(dir .eq. 1) then dok=1,NZ doj=1,NY 109 do i = 1, NX a(m) = tempa(i,j,k) b(m) = tempb(i,j,k) c(m) = tempc(i,j,k) =m+1 enddo enddo enddo elseif(dir .eq. 2) then do k = 1, NZ do i = 1, NX do j = 1, NY a(m) = tempa(i,j,k) b(m) = tempb(i,j,k) c(m) = tempc(i,j,k) m=m+l enddo enddo enddo endif return END ************** subroutine flipr(r,dir) this subroutine reorders the r vector for the tridag routine to work properly. CHANGE model size here integer NMAX,NX,NY,NZ parameter (NX = 41) parameter (NY = 41) parameter (NZ = 41) parameter (NMAX = 68921) integer i,j,k,m,dir real(8) r(NMAX) real(8) tempr(NX,NY,NZ) 110 l h, :5 :31: gym-"J m = 1 call ianos(r,tempr,3) if(dir .eq. 1) then do k = 1, NZ doj = 1, NY do i = l, NX r(m) = tempr(i,j,k) m=m+1 enddo enddo enddo elseif(dir .eq. 2) then do k = 1, NZ do i = l, NX doj = 1, NY r(m) = tempr(i,j,k) m=m+1 enddo enddo enddo endif return END ************** subroutine invpos(in,out,dir) this transforms the vector that is outputted by tridag into an array. Should save on function calls. CHANGE model size here integer NMAX,NX,NY,NZ parameter (NX = 41) parameter (NY = 41) parameter (NZ = 41) parameter (NMAX = 68921) 111 integer i,j,k,LOCCOUNT,dir real(8) in(NMAX),out(NX,NY,NZ) LOCCOUNT = 0 if(dir.eq. 1) then do k = 1, NZ do j = 1, NY do i = l, NX LOCCOUNT = LOCCOUNT + 1 out(i,j,k) = in(LOCCOUNT) enddo enddo enddo elseif(dir.eq.2) then do k = 1, NZ do i = 1, NX do j = 1, NY LOCCOUNT = LOCCOUNT + 1 out(i,j,k) = in(LOCCOUNT) enddo enddo enddo elseif(dir.eq.3) then do i = l, NX do j = 1, NY do k = 1, NZ LOCCOUNT = LOCCOUNT + l out(i,j,k) = in(LOCCOUNT) enddo enddo enddo endif return END **************************** subroutine output(count) This routine sucks the points of interest from the temperature 112 0 0 ()(5 0 vector (T) at each time step and saves it to a handy vector to send back to the main program. This is what gets compared to the experimental data for the Marquardt optimization. write as an array here, output to file at end of routine CHANGE model size here integer NMAX,NX,NY,NZ parameter (NX = 41) parameter (NY = 41) parameter (NZ = 41) parameter (NMAX = 68921) integer x(5),y(5),z(5),count,XTC1,YTC1,XTC2,YTC2 real(8) T(NX,NY,NZ), Tout(50000,5),xcorr,ycorr real(8) tone(NX,NY,NZ),ttwo(NX,NY,NZ) common IT EMPERATURE/ T,tone,ttwo,Tout common IOTPT/ x,y,z,XTC 1 ,XTC2,YTC l ,YTC2,xcorr,ycorr Tout(count,1) = T(x(1), y(l), z(1)) Tout(count,2) = T(x(2), y(2), 2(2)) Tout(count,3) = xcorr*T(XTC1,y(3),z(3)) + + (l .-xcorr)*T(XTC2,y(3),z(3)) Tout(count,4) = ycorr*T(x(4),YTCl,z(4)) + + (l .-ycorr)*T(x(4),YTC2,z(4)) Tout(count,5) = T(x(S), y(5), 2(5)) return END ************** subroutine tridag(a,b,c,r,u,n) Thomas Algorithm for solving a tridiagonal matrix. This is culled from Numerical Recipes in FORTRAN. Solves for a vector u(1:n) of length n a tridiagonal linear set. a(l :n), b(l :n), c(1:n), and r(l :n) are input vectors and are not modified. Parameter: NMAX is the maximum expected value of n. integer N MAX parameter (NMAX = 100000) 113 integer n real(8) a(n),b(n),c(n),r(n),u(n) integer j real(8) bet,gam(NMAX) if(b(l).eq.0.)pause 'tridag: rewrite equations' bet=b( 1 ) u(1 )=r(1 )/bet do j=2,n gam(i)=c(j-1 )/bet bet=b0)-a0)*gam0) if(bet.eq.0.)pause 'tridag failed (bet=0)’ u0)=(r0)-a0)*uo-1»/bet enddo do j=n-l,1,-l ud)=u0)-gamo+1)*u0+1> enddo return END C ***************** 0000 00 subroutine rfill(r,STEP,fs) This routine will fill the r-vector for use in tridag. It will be called once during each step. Outputs the vector r, and increments the STEP count by one. This takes the bulk of the program time, I think, and may be worth optimizing. NOTE: going from 3-space to vector space would save computational time, but would also be a brain-buster. CHANGE model size here integer NMAX,NX,NY,NZ parameter (NX = 41) parameter (NY = 41) parameter (NZ = 41) parameter (NMAX = 68921) integer i,j,k,STEP,position,QI,QJ real(8) r(NMAX),Tout(50000,5) real(8) T(NX,NY,NZ),tone(NX,NY,NZ) real(8) ttwo(NX,NY,NZ) real(8) Fo,Gl ,G2,Biot,q,Bind,qnd,nBind,dx,dy,dz 114 real(8) fs,QIB,QJB,bBind,anind real(8) RA(9),RB(9),RC(9),RD(9),RE(9),RF(9),RG(9) real(8) RH(9),RI(9),RJ(9),RK(9),RL(9),RM(9),RN(9) real(8) RO(9),RP(9),RQ(9),RR(9) common ITEMPERATURE/ T,tone,ttwo,Tout common lRVECT/ RA,RB,RC,RD,RE,RP,RG,RH,RI,RJ,RK,RL,RM,RN,RO common /RVECT2/ RP,RQ,RR,bBind,anind common IPROPS/ Fo,G1,G2,Biot,q,qnd,Bind,nBind common /RFU QI,QJ,QIB,QJB common /GEOM/ dx,dy,dz internal computations do i = 2, NX-l do j = 2, NY-l do k = 2, NZ-l r(position(i,j,k))=RA(STEP)*T(i,j,k)+RB(STEP)* (T(i+1,j,k)+T(i-1,j,k))+RC(STEP)*(T(i,j+1,k)+ T(i,j- l,k))+RD(STEP)*(T(i,j,k+1)+T(i,j,k-l ))+ RE(STEP)*tone(i,j,k)+RF(STEP)*(tone(i+1 ,j,k)+tone (i-1,j,k))+RG(STEP)*(tone(i,j+1,k)+tone(i,j-1,k))+ RH(STEP)*(tone(i, j,k+1 )+tone(i, j ,k- 1 ))+RI(STEP)* ttwo(i,j,k)+RJ(STEP)*(ttwo(i+1,j,k)+ttwo(i-1,j,k)) +RK(STEP)*(ttwo(i,j+1 ,k)+ttwo(i,j- l ,k))+RL(STEP)* (ttwo(i,j,k+1)+ttwo(i,j,k-1)) enddo ++++++++ 2 face computations r(position(i,j, 1 ))=RA(STEP)*T(i,j, l )+RB(STEP)* (T(i+1 ,j,l)+T(i-1,j,1))+RC(STEP)*(T(i,j+l,l)+ T(i,j-l,1))+RD(STEP)*(2.*T(i,j,2))+ RE(STEP)*tone(i, j, l )+RF(STEP)*(tone(i+1,j,1)+tone (i-l,j,1))+RG(STEP)*(tone(i,j+1 ,l )+tone(i,j-1,1 ))+ RH(STEP)*(2.*tone(i,j,2))+RI(STEP)* ttwo(i,j,1)+RJ(STEP)*(ttwo(i+1,j,1)+ttwo(i-1,j,] )) +RK(STEP)*(ttwo(i,j+l ,1 )+ttwo(i,j-1 ,1 ))+RL(STEP)* (2.*ttwo(i,j,2))+RM(STEP)*(T(i,j,1))+ RN(STEP)*(tone(i,j,1))+RO(STEP)*(ttwo(i,j,1)) +++++++++ r(position(i,j,NZ)):RA(STEP)*T(i,j,NZ)+RB(STEP)* (T(i+1,j,NZ)+T(i-1,j,NZ))+RC(STEP)*(T(i,j+l ,NZ)+ T(i,j-l,NZ))+RD(STEP)*(2.*T(i,j,NZ-1))+ RE(STEP)*tone(i,j,NZ)+RF(STEP)*(tone(i+1 ,j,NZ)+tone (i-l,j,NZ))+RG(STEP)*(tone(i,j+1,NZ)+tone(i,j-1,NZ))+ RH(STEP)*(2.*tone(i,j,NZ-1))+RI(STEP)* +++++ 115 ++++ ttwo(i,j,NZ)+RJ(STEP)*(ttwo(i+1 ,j,NZ)+ttwo(i-1 ,j,NZ)) +RK(STEP)*(ttwo(i,j+l,NZ)+ttwo(i,j-1,NZ))+RL(STEP)* (2.*ttwo(i,j,NZ-1))+RP(STEP)*(T(i,j,NZ))+ RQ(STEP)*(tone(i,j,NZ))+RR(STEP)*(ttwo(i,j,NZ)) enddo compute four edges +++++++++ +++++++++ +++++++++ ++++++ r(position(i,1,1))=RA(STEP)*T(i,1,1)+RB(STEP)* (T(i+1,1,1)+T(i-1,1,1))+RC(STEP)*(2.*T(i,2,1))-1- RD(STEP)*(2.*T(i,1,2))+ RE(STEP)*tone(i,l,1)+RF(STEP)*(tone(i+1,1 ,1 )+tone (i-1,1,1))+RG(STEP)*(2.*tone(i,2,1))+ RH(STEP)*(2.*tone(i, 1 ,2))+RI(STEP)* ttwo(i,1,1)+RJ(STEP)*(ttwo(i+1,1,1)+ttwo(i-1 ,1 ,1 )) +RK(STEP)*(2.*ttwo(i,2,1))+RL(STEP)* (2.*ttwo(i,1,2))+RM(STEP)*(T(i,1,1))+ RN(STEP)*(tone(i,],1))+RO(STEP)*(ttwo(i,1,1)) r(position(i, 1 ,NZ))=RA(STEP)*T(i, l ,NZ)+RB(STEP)* (T(i+l,l ,NZ)+T(i-1,1,NZ))+RC(STEP)*(2.*T(i,2,NZ)) +RD(STEP)*(2.*T(i,1,NZ-l ))+ RE(STEP)*tone(i,1 ,NZ)+RF(STEP)*(tone(i+1 ,1 ,NZ)+tone (i-1,1,NZ))+RG(STEP)*(2.*tone(i,2,NZ))+ RH(STEP)*(2.*tone(i,l,NZ-1))+RI(STEP)* ttwo(i,],NZ)+RJ(STEP)*(ttwo(i+1,1,NZ)+ttwo(i-1 ,1 ,NZ)) +RK(STEP)*(2.*ttwo(i,2,NZ))+RL(STEP)* (2.*ttwo(i,1,NZ—1))+RP(STEP)*(T(i,l ,NZ))+ RQ(STEP)*(tone(i,1 ,NZ))+RR(STEP)*(ttwo(i,1 ,NZ)) r(position(i,NY,l ))=RA(STEP)*T(i,NY,1)+RB(STEP)* (T(i+l ,NY,] )+T(i-1,NY,1))+RC(STEP)* (2.*T(i,NY-1,1 ))+RD(STEP)*(2.*T(i,NY,2))+ RE(STEP)*tone(i,NY,1)+RF(STEP)*(tone(i+l ,NY,] )+tone (i-1,NY,]))+RG(STEP)*(2.*tone(i,NY-1,l))+ RH(STEP)*(2.*tone(i,NY,2))+RI(STEP)* ttwo(i,NY,])+RJ(STEP)*(ttwo(i+1,NY,1)+ttwo(i-1,NY,1)) +RK(STEP)*(2.*ttwo(i,NY-1,1))+RL(STEP)* (2.*ttwo(i,NY,2))+RM(STEP)*(T(i,N Y, 1 ))+ RN(STEP)*(tone(i,NY,1))+RO(STEP)*(ttwo(i,NY,1)) r(position(i,NY,NZ))=RA(STEP)*T(i,NY,NZ)+RB(STEP)* (T(i+l ,NY,NZ)+T(i- l ,NY,NZ))+RC(STEP)* (2. *T(i,NY- 1 ,NZ))+RD(STEP)*(2.*T(i,NY,NZ- 1 ))+ RE(STEP)*tone(i,NY,NZ)+RF(STEP)*(tone(i+1,NY,NZ)+tone (i-l,NY,NZ))+RG(STEP)*(2.*tone(i,NY-1,NZ))+ RH(STEP)*(2.*tone(i,NY,NZ—1))+RI(STEP)* ttwo(i,NY,NZ)+RJ(STEP)*(ttwo(i+1,NY,NZ)+ttwo(i-1,NY,NZ)) 116 + + + enddo +RK(STEP)*(2.*ttwo(i,NY-1,NZ))+RL(STEP)* (2.*ttwo(i,NY,NZ-1 ))+RP(STEP)*(T(1,NY,NZ))-I- RQ(STEP)*(tone(i,NY,NZ))+RR(STEP)*(ttwo(i,NY,NZ)) two more faces doj =2, NY-l ++++++++ ++++++++ do k = 2, NZ—l r(position( 1 ,j,k))=RA(STEP)*T( 1 ,j,k)+RB(STEP)* (2.*T(2,j,k))+RC(STEP)*(T( 1 ,j+1 ,k)+ T(1,j-1,k))+RD(STEP)*(T(1 ,j,k+1)+T(1,j,k-1))+ RE(STEP)*tone( 1 ,j,k)+RF(STEP)*(2.*tone(2,j,k)) +RG(STEP)*(tone(l,j+1,k)+tone(1,j-1,k))+ RH(STEP)*(tone(1,j,k+l)+tone(1 ,j,k- l ))+RI(STEP)* ttwo(l,j,k)+RJ(STEP)*(2.*ttwo(2,j,k)) +RK(STEP)*(ttwo(1,j+1,k)+ttwo(1 ,j- 1 ,k))+RL(STEP)* (ttwo(l ,j,k+l)+ttwo(1,j,k-1 )) r(position(NX,j,k))=RA(STEP)*T(NX,j,k)+RB(STEP)* (2.*T(NX-l,j,k))+RC(STEP)*(T(NX,j+1,k)+ T(NX,j-1,k))+RD(STEP)*(T(NX,j,k+1)+T(NX,j,k-1))+ RE(STEP)*tone(NX,j,k)+RF(STEP)*(2.*tone (NX-1,j,k))+RG(STEP)*(tone(NX,j+l ,k)+tone(NX,j-l,k))+ RH(STEP)*(tone(NX,j,k+1)+tone(NX,j,k-l ))+RI(STEP)* ttwo(NX,j,k)+RJ(STEP)*(2.*ttwo(NX-1,j,k)) +RK(STEP)*(ttwo(NX,j+1 ,k)+ttwo(NX,j-1 ,k))+RL(STEP)* (ttwo(NX,j,k+1)+ttwo(NX,j,k-l )) enddo four more edges +++++++++ + +-+ + r(position(l,j,1))=RA(STEP)*T(1,j,1)+RB(STEP)* (2.*T(2,j,1 ))+RC(STEP)*(T(1 ,j+1,1)+ T(1,j-1,l))+RD(STEP)*(2.*T(1,j,2))+ RE(STEP)*tone(l,j,1)+RF(STEP)*(2.*tone(2,j,1)) +RG(STEP)*(tone(l,j+l,1)+tone(1,j-1,1))+ RH(STEP)*(2.*tone(1 ,j,2))+RI(STEP)* ttwo(l,j,1)+RJ(STEP)*(2.*ttwo(2,j,1)) +RK(STEP)*(ttwo(1 ,j+1 ,1)+ttwo(l ,j-l ,1 ))+RL(STEP)* (2.*ttwo(1,j,2))+RM(STEP)*(T(1,j,1))+ RN(STEP)*(tone(l,j,1))+RO(STEP)*(ttwo(1,j,1)) r(position(],j,NZ))=RA(STEP)*T(1,j,NZ)+RB(STEP)* (2.*T(2,j,NZ))+RC(STEP)*(T(1,j+1 ,NZ)+ T(1,j-1,NZ))+RD(STEP)*(2.*T(1,j,NZ-1))+ RE(STEP)*tone(1,j,NZ)+RF(STEP)*(2.*tone(2,j,NZ)) +RG(STEP)*(tone(1,j+1,NZ)+tone(1,j-1,NZ))+ 117 +++++ +++++++++ +++++++++ 0 :3 D... D. O RH(STEP)*(2.*tone(1,j,NZ-l))+RI(STEP)* ttwo(l ,j,NZ)+RJ(STEP)*(2.*ttwo(2,j,NZ)) +RK(STEP)*(ttwo(l,j+l,NZ)+ttwo(1,j-1,NZ))+RL(STEP)* (2.*ttwo(1,j,NZ-1))+RP(STEP)*(T(1,j,NZ))+ RQ(STEP)*(tone(1,j,NZ))+RR(STEP)*(ttwo(1,j,NZ)) r(position(NX, j, 1 ))=RA(STEP)*T(N X,j, l )+RB(STEP)* (2.*T(NX-l,j,l))+RC(STEP)*(T(NX,j+1,1)+ T(NX,j-1,1))+RD(STEP)*(2.*T(NX,j,2))+ RE(STEP)*tone(NX,j,1)+RF(STEP)* (2.*tone(NX-1,j,1))+RG(STEP)*(tone(NX,j+1,l)+tone(NX,j-1 ,1 ))+ RH(STEP)*(2.*tone(NX,j,2))+RI(STEP)* ttwo(NX,j,])+RJ(STEP)*(2.*ttwo(NX-1,j,1)) +RK(STEP)*(ttwo(NX,j+l ,1 )+ttwo(NX,j-1,1 ))+RL(STEP)* (2.*ttwo(NX,j,2))+RM(STEP)*(T(NX,j,1))+ RN (STEP)*(tone(NX,j, 1 ))+RO(STEP)*(ttwo(N X,j, l )) r(position(NX,j,NZ))=RA(STEP)*T(NX,j,NZ)+RB(STEP)* (2.*T(NX-1,j,NZ))+RC(STEP)*(T(NX,j+1,NZ)+ T(NX,j-l,NZ))+RD(STEP)*(2.*T(NX,j,NZ-l))+ RE(STEP)*tone(NX,j,NZ)+RF(STEP)* (2.*tone(NX-1,j,NZ))+RG(STEP)*(tone(NX,j+1,NZ)+ tone(NX,j-1,NZ))+RH(STEP)*(2.*tone(NX,j,NZ-1))+RI(STEP)* ttwo(NX,j,NZ)+RJ(STEP)*(2.*ttwo(NX-l,j,NZ)) +RK(STEP)*(ttwo(NX,j+1,NZ)+ttwo(NX,j—l,NZ))+RL(STEP)* (2.*ttwo(NX,j,NZ-1))+RP(STEP)*(T(NX,j,NZ))+ RQ(STEP)*(tone(NX,j,NZ))+RR(STEP)*(ttwo(NX,j,NZ)) last two faces do k = 2, NZ-l do i = 2, NX-l r(position(i, 1 ,k))=RA(STEP)*T(i, 1 ,k)+RB(STEP)* + (T(i+l,1,k)+T(i-1,l,k))+RC(STEP)*(2.*T(i,2,k)) + +RD(STEP)*(T(i,1,k+l)+T(i,1,k-1))+ + RE(STEP)*tone(i, 1 ,k)+RF(STEP)*(tone(i+1 ,1 ,k)+-tone + (i- 1 , l ,k))+RG(STEP)*(2.*tone(i,2,k))+ + RH(STEP)*(tone(i,l,k+l)+tone(i,1,k-1))+RI(STEP)* + ttwo(i,1,k)+RJ(STEP)*(ttwo(i+1 ,l,k)+ttwo(i-l,1,k)) + +RK(STEP)*(2.*ttwo(i,2,k))+RL(STEP)* + (ttwo(i,l,k+1)+ttwo(i,1,k-1)) r(position(i,NY,k))=RA(STEP)*T(i,NY,k)+RB(STEP)* + (T(i+1,NY,k)+T(i-1,NY,k))+RC(STEP)*(2.* + T(i,NY-l,k))+RD(STEP)*(T(i,NY,k+1)+T(i,NY,k-1))+ 118 ++++++ RE(STEP)*tone(i,NY,k)+RF(STEP)*(tone(i+1,NY,k)+tone (i-1,NY,k))+RG(STEP)*(2.*tone(i,NY-l ,k))+ RH(STEP)*(tone(i,NY,k+1)+tone(i,NY,k-1))+RI(STEP)* ttwo(i,NY,k)+RJ(STEP)*(ttwo(i+1,NY,k)+ttwo(i-1,NY,k)) +RK(STEP)*(2.*ttwo(i,NY-1,k))+RL(STEP)* (ttwo(i,NY,k+1)+ttwo(i,NY,k-1)) enddo last four edges ++++++++ ++++++++ ++++++++ ++++++ r(position(l,l,k))=RA(STEP)*T(1,1,k)+RB(STEP)* (2.*T(2,1,k))+RC(STEP)*(2.*T(1,2,k))+ RD(STEP)*(T(] ,1 ,k+1 )+T(] ,1,k-1 ))+ RE(STEP)*tone(1,1,k)+RF(STEP)*(2.*tone(2,1,k)) +RG(STEP)*(2.*tone(1,2,k))+ RH(STEP)*(tone(],1,k+1)+tone(1,1,k-l))+RI(STEP)* ttwo(1,1,k)+RJ(STEP)*(2.*ttwo(2,1,k)) +RK(STEP)*(2.*ttwo(1,2,k))+RL(STEP)* (ttwo(1,1,k+1)+ttwo(l ,1,k-1)) r(position( 1 ,NY,k))=RA(STEP)*T( l ,NY,k)+RB(STEP)* (2.*T(2,NY,k))+RC(STEP)* (2.*T(1,NY-1,k))+RD(STEP)*(T(1,NY,k+l)+T(1 ,NY,k-1))+ RE(STEP)*tone( l ,NY,k)+RF(STEP)*(2. *tone(2,NY,k)) +RG(STEP)*(2.*tone(1,NY-1,k))+ RH(STEP)*(tone(] ,NY,k+1)+tone(1 ,NY,k-l ))+RI(STEP)* ttwo( 1 ,NY,k)+RJ(STEP)*(2. *ttwo(2,NY,k)) +RK(STEP)*(2.*ttwo(1,NY-l ,k))+RL(STEP)* (ttwo(l,NY,k+1)+ttwo(l,NY,k-1)) r(position(NX, 1 ,k))=RA(STEP)*T(NX, 1 ,k)+RB(STEP)* (2.*T(NX-1,l,k))+RC(STEP)*(2.*T(NX,2,k)) +RD(STEP)*(T(N X, 1 ,k-t- 1 )+T(NX, l ,k- l ))+ RE(STEP)*tone(NX, 1 ,k)+RF(STEP)* (2.*tone(NX-1,1,k))+RG(STEP)*(2.*tone(NX,2,k))+ RH(STEP)*(tone(NX,1,k+1)+tone(NX, l ,k-l ))+RI(STEP)* ttwo(NX, 1 ,k)+RJ(STEP)*(2.*ttwo(NX-l ,1 ,k)) +RK(STEP)*(2.*ttwo(NX,2,k))+RL(STEP)* (ttwo(N X, l ,k+1 )+ttwo(NX, l ,k-1)) r(position(NX,NY,k))=RA(STEP)*T(NX,NY,k)+RB(STEP)* (2.*T(NX-1,NY,k))+RC(STEP)*(2.*T(NX,NY-1,k)) +RD(STEP)*(T(NX, NY ,k+1)+T(NX, NY, k- 1))+ RE(STEP)*tone(NX, NY ,.k)+RF(STEP)*(2 *tone(NX- 1 ,NY ,k)) +RG(STEP)*(2.*tone(NX,NY-1,k))+ RH(STEP)*(tone(NX, NY, k+1 )+tone(NX,NY,k- 1 ))+RI(STEP)* ttwo(NX,NY,k)+RJ(STEP)*(2.*ttwo(NX- l ,NY,k)) 119 + + enddo +RK(STEP)*(2. *ttwo(NX,NY- l ,k))+RL(STEP)* (ttwo(NX,NY,k+1)+ttwo(NX,NY,k-1)) compute the eight corners +++++++++ +++++++++ +++++++++ ++++++ r(position(],1,1))=RA(STEP)*T(1,1,l )+RB(STEP)* (2.*T(2,1,1))+RC(STEP)*(2.*T(1,2,1))+ RD(STEP)*(2.*T(1,1,2))+ RE(STEP)*tone(],l,1)+RF(STEP)*(2.*tone(2,1,1)) +RG(STEP)*(2.*tone(l ,2,1))+ RH(STEP)*(2.*tone( 1,1,2))+RI(STEP)* ttwo(l,1,1)+RJ(STEP)*(2.*ttwo(2,1,1)) +RK(STEP)*(2.*ttwo(1,2,1))+RL(STEP)* (2.*ttwo(1,1,2))+RM(STEP)*(T(1,1,1))+ RN(STEP)*(tone(],l,1))+RO(STEP)*(ttwo(1,1,1)) r(position(l,l ,NZ))=RA(STEP)*T(1,1,NZ)+RB(STEP)* (2.*T(2,1,NZ))+RC(STEP)*(2.*T(1,2,NZ)) +RD(STEP)*(2.*T(1,1,NZ-1))+ RE(STEP)*tone(],1,NZ)+RF(STEP)*(2.*tone(2,1,NZ))+ RG(STEP)*(2.*tone(1,2,NZ))+ RH(STEP)*(2.*tone(] ,1,NZ-1))+RI(STEP)* ttwo(l,1,NZ)+RJ(STEP)*(2.*ttwo(2,1,NZ)) +RK(STEP)*(2.*ttwo(1 ,2,NZ))+RL(STEP)* (2.*ttwo(l,l,NZ-1))+RP(STEP)*(T(1 ,l ,NZ))+ RQ(STEP)*(tone( l , 1 ,NZ))+RR(STEP)*(ttwo(1 ,1 ,NZ)) r(position(],NY,]))=RA(STEP)*T(1,NY,1)+RB(STEP)* (2.*T(2,NY,1))+RC(STEP)* (2.*T(1,NY-1,1))+RD(STEP)*(2.*T(1,NY,2))+ RE(STEP)*tone( 1 ,NY, 1 )+RF(STEP)*(2. *tone(2,NY, 1 )) +RG(STEP)*(2.*tone(1,NY-1,1))+ RH(STEP)*(2.*tone( 1 ,NY,2))+RI(STEP)* ttwo(1,NY,1 )+RJ(STEP)*(2.*ttwo(2,NY,1)) +RK(STEP)*(2.*ttwo(1 ,NY-1,1 ))+RL(STEP)* (2.*ttwo(1,NY,2))+RM(STEP)*(T(1,NY,1))+ RN (STEP)*(tone( 1 ,NY, 1 ))+RO(STEP)*(ttwo( 1 ,N Y, 1 )) r(position(] ,NY,NZ))=RA(STEP)*T(1 ,NY,NZ)+RB(STEP)* (2.*T(2,NY,NZ))+RC(STEP)* (2.*T(l,NY-1,NZ))+RD(STEP)*(2.*T(1,NY,NZ—l))+ RE(STEP)*tone( 1 ,NY,NZ)+RF(STEP)*(2.*tone(2,NY,NZ)) +RG(STEP)*(2.*tone(1,NY-1,NZ))+ RH(STEP)*(2.*tone(1,NY,NZ-1))+RI(STEP)* ttwo( 1 ,NY,NZ)+RJ(STEP)*(2.*ttwo(2,NY,NZ)) 120 ++ +++++++++ +++++++++ +++++++++ ++++++++ +RK(STEP)*(2.*ttwo(1 ,NY-1 ,NZ))+RL(STEP)* (2.*ttwo(l ,NY,NZ-l ))+RP(STEP)*(T(1,NY,NZ))+ RQ(STEP)*(t0ne(1,NY,NZ))+RR(STEP)*(ttwo( 1 ,NY,NZ)) r(position(NX,],l))=RA(STEP)*T(NX,1,1)+RB(STEP)* (2.*T(NX-1,l,1))+RC(STEP)*(2.*T(NX,2,1)) +RD(STEP)*(2.*T(NX,1,2))+ RE(STEP)*tone(NX,1,1)+RF(STEP)*(2.*tone (NX-1 ,1 ,1))+RG(STEP)*(2.*tone(NX,2,1))+ RH(STEP)*(2.*tone(NX, l ,2))+RI(STEP)* ttwo(NX,l,1)+RJ(STEP)*(2.*ttwo(NX-1,l,1)) +RK(STEP)*(2.*ttwo(NX,2,1))+RL(STEP)* (2.*ttwo(NX,l ,2))+RM(STEP)*(T(NX,1,1))+ RN (STEP)*(tone(NX, 1 ,1))+RO(STEP)*(ttwo(NX,1,1)) r(position(NX,1,NZ))=RA(STEP)*T(NX,l ,NZ)+RB(STEP)* (2. *T(NX- l , l ,NZ))+RC(STEP)*(2.*T(NX,2,NZ)) +RD(STEP)*(2.*T(N X, 1 ,NZ- 1 ))+ RE(STEP)*tone(N X, 1 ,NZ)+RF(STEP)*(2. *tone (NX-l,1,NZ))+RG(STEP)*(2.*tone(NX,2,NZ))+ RH(STEP)*(2.*tone(N X, 1 ,NZ- 1 ))+RI(STEP)* ttwo(NX, 1 ,NZ)+RJ(STEP)*(2.*ttwo(NX-1,1 ,NZ)) +RK(STEP)*(2.*ttwo(NX,2,NZ))+RL(STEP)* (2.*ttwo(N X, 1 ,NZ- 1 ))+RP(STEP)*(T(NX, l ,N Z))+ RQ(STEP)*(tone(N X, 1 ,NZ))+RR(STEP)*(ttwo(NX, 1 ,NZ)) r(position(NX,NY,1))=RA(STEP)*T(NX,NY,1)+RB(STEP)* (2.*T(NX-1,NY,1))+RC(STEP)* (2.*T(NX,NY-1,l))+RD(STEP)*(2.*T(NX,NY,2))+ RE(STEP)*tone(NX,NY, 1 )+RF(STEP)*(2. *tone (NX-1,NY,]))+RG(STEP)*(2.*tone(NX,NY-1,1))+ RH(STEP)*(2.*tone(NX,NY,2))+RI(STEP)* ttwo(NX,NY, 1 )+RJ (STEP)*(2. *ttwo(NX- 1 ,N Y, 1 )) +RK(STEP)*(2.*ttwo(NX,NY-l ,1))+RL(STEP)* (2. *ttwo(NX,N Y,2))+RM(STEP)*(T(NX,N Y, 1 ))+ RN (STEP)*(tone(NX,N Y, 1 ))+RO(STEP)*(ttwo(N X,NY, 1 )) r(position(NX,NY,NZ))=RA(STEP)*T(NX,NY,NZ)+RB(STEP)* (2.*T(NX-1,NY,NZ))+RC(STEP)* (2.*T(NX,NY-1 ,NZ))+RD(STEP)*(2.*T(NX,NY,NZ-l ))+ RE(STEP)*tone(NX,NY,NZ)+RF(STEP)*(2.*tone (NX-1,NY,NZ))+RG(STEP)*(2.*tone(NX,NY-1,NZ))+ RH(STEP)*(Z.*tone(NX,NY,NZ-1))+RI(STEP)* ttwo(NX,NY,NZ)+RJ(STEP)*(2.*ttwo(NX-1,NY,NZ)) +RK(STEP)*(2.*ttwo(NX,NY-1,NZ))+RL(STEP)* (2.*ttwo(NX,NY,NZ-1))+RP(STEP)*(T(NX,NY,NZ))+ 121 0 (If? 0 (DC? 0 (>15 0 ()(7 0 ()(1 0 ()(5 0 ()(fi 0 ()(5 0 ()(5 0 (3(3 0 ()(3(1 0 + RQ(STEP)*(tone(NX,NY,NZ))+RR(STEP)*(ttwo(NX,NY,NZ)) FLUX SECTION -- here, the illuminated area is set do i = 1, Q1 do j = 1, Q] r(position(i,j, 1)) = r(position(i,j, l ))+fs*qnd enddo enddo *********************** DIFFERENT APERTURE WIDTHS NOTE: this is set for a 1 1.14mm width square beam with comers that have a ~3.17mm fillet. (5.57 xapt) these terms add a portion of the total flux to the boundary of the heated surface -- needed if the actual geometry doesn't line up with the descretized geometry. borderline edge nodes -- x-edge do j = 1, 7 do j = 1, Q] r(position(QI+l ,j,1)) = r(position(QI+1,j,1))+fs*QIB *qnd enddo borderline edge nodes -- y-edge do i = l, 7 do i = 1, Q1 r(position(i,QJ +1 , l )) = r(position(i,QJ+1,1))+fs*QJB*qnd enddo borderline comer node r(position(QI+1 ,QJ +1 , 1 )) = r(position(QI+l,QJ+1,1))+fs*QIB *QJ B *qnd subtract out the interior nodes that are not to be heated r(position(l 1,10,1)) = r(position(l 1,10,1))-fs*qnd r(position( 1 1,1 1,1 )) = r(position(] 1,1 1,1 ))-fs*qnd r(position(10,1 1.1)) = r(position( 10,1 1,1))-fs*qnd ************************* this section for an 11.60mm width beam with comers that have a ~3.17mm fillet XAPT = 5.85 122 0 ()(5 0 ()(5 0 ()(5 0 ()(5 0 (>(5 0 ()(1 0 (3(7 0 borderline edge nodes -- x-edge do j = l, 6 r(position(QI+1,j,l )) = r(position(QI+1,j,1))+fs*QIB *qnd enddo borderline edge nodes -- y-edge do i = 1, 6 r(position(i,QJ +1 , 1 )) = r(position(i,QJ+],1))+fs*QJB*qnd enddo subtract out the interior nodes that are not to be heated r(position(12,10,1)) = r(position(12,10,l ))-fs*qnd r(position(12,1 l ,1)) = r(position(12,1 1,1))-fs*qnd r(position(12,12,l )) = r(position(12,12,1))-fs*qnd r(position(10,12,1)) = r(position(10,12,1))-fs*qnd r(position(] 1,12,1)) = r(position(l 1,12,1))-fs*qnd r(position(] 1,1 1,1)) = r(position(] 1,1 1,1))-0.5*fs*qnd r(position(12,9,l)) = r(position(12,9,1))-0.35*fs*qnd r(position(9,] 2,1)) = r(position(9,]2,1))-0.35*fs*qnd ******************** this section for an 10.62mm width beam with corners that have a ~3.17mm fillet (5.31 xapt) borderline edge nodes -- x-edge do j = 1, 4 r(position(QI+1,j,l )) = r(position(QI+1,j,l))+fs*QIB *qnd enddo borderline edge nodes -- y-edge do i = 1, 4 r(position(i,QJ +1 , l )) = r(position(i,QJ+1,1))+fs*QJB*qnd enddo subtract out the interior nodes that are not to be heated r(position(] 1,10,1)) = r(position(l l ,10,1))-fs*qnd r(position(] 1,1 1, l )) = r(position(] 1,1 1 ,1))-fs*qnd r(position(10,l 1,1)) = r(position(10,1 1,1))-fs*qnd r(position(9,] 1,1)) = r(position(9,1 1,1 ))-fs*qnd r(position(] 1 ,9,1)) = r(position(] 1,9,1))-fs*qnd 123 0 (DC? 0 (DC? 0 ()(5 0 (DC? 0 (it? 0 (3(3 0 ()(5 0 (3!? 0 ()(7 0 ()(7 0 ()(1 0 ()(7 0 (if? 0 ()(5 0 ()(3 0 r(position(10,10, 1)) = r(position(10,10,1))-fs*qnd r(position(8,] 1,1)) = r(position(8,1 1,1))—0.5 *fs*qnd r(position(] 1,8,1 )) = r(position(] l ,8,1))-0.5*fs*qnd **************************** NOTE: this is set for an 11.30mm width square beam with comers that have a ~3.17mm fillet. XAPT = 5.65 these terms add a portion of the total flux to the boundary of the heated surface -- needed if the actual geometry doesn't line up with the descretized geometry. borderline edge nodes -- x-edge do j = l, 6 r(position(QI+l,j,1)) = r(position(QI+1,j,1))+fs*QIB *qnd enddo borderline edge nodes -- y-edge do i = 1, 6 r(position(i,QJ +1 , 1 )) = r(position(i,QJ +1 , l ))+fs*QJ B*qnd enddo add nodes on edge that are partially heated r(position(12,9,l)) = r(position(12,9,1))+0.1*fs*QIB *qnd r(position(9,]2,1)) = r(position(9,]2,1))+0.1*fs*QJB*qnd r(position(12,8,l)) = r(position(12,8,1))+0.55 *fs*QIB*qnd r(position(8,]2,1)) = r(position(8, 12,1 ))+0.55*fs*QJB*qnd r(position(12,7,1)) = r(position(12,7,l ))+0.95*fs*QIB*qnd r(position(7,12, 1 )) = r(position(7,12,1))+0.95*fs*QJB *qnd subtract out the interior nodes that are not to be heated r(position(] 1,1 1,1)) = r(position(l 1,1 1, l ))-fs*qnd r(position(] 1,10,1)) = r(position(] 1,10, l ))-0.8*fs*qnd r(position(10,1 1, l )) = r(position(10,1 1,1))-0.8*fs*qnd ******************** OLD ALTER APERTURE SIZE CODE this is the section where nodes are suntracted off (if not merely commented out above). i=12 do i = 1 1,12 r(position(i, 1 3,1)) = r(position(i,13,1))-fs*QIB*qnd 124 0 (DC? 0 ()(5 0 ()(3 0 ()(5 0 ()(3 0 (>(5 0 ()(5 0 (ht? 0 (3!? 0 (3(5 0 0 r(position(13,i,1)) = r(position(13,i,1 ))-fs*QJB*qnd enddo r(position(12,12,1)) = r(position(12,12,1))-fs*qnd EDWIN-LIKE APERTURE CODE fix convection to make it like Edwin's model do i = 1, 6 do j = 1, 6 r(position(i,j,1)) = r(position(i,j,1)) + -fs*RM(STEP)*(T(i,j,1))—fs*RN(STEP)*(tone(i,j,1)) -fs*RO(STEP)*(ttwo(i,j, l )) enddo enddo do i = 1, 6 r(position(i,7,1)) = r(position(i,7, l ))-fs*0.65 *RM(STEP)* + (T(i,7,1))-fs*0.65*RN(STEP)*(tone(i,7,1))-fs*0.65*RO(STEP)* (ttwo(i,7, l )) enddo do j = l, 6 r(position(7,j,1)) = r(position(7,j, 1 ))-fs*0.65 *RM(STEP)* + (T(7,j,1))-fs*0.65 *RN(STEP)*(tone(7,j,1))-fs*0.65*RO(STEP)* (ttwo(7,j,1)) enddo r(position(7,7,1)) = r(position(7,7,1))-fs*0.65 *0.65 *RM(STEP)* +(T(7,7,1))-fs*0.65 *0.65 *RN(STEP)*(tone(7,7, l ))-fs*0.65*0.65* +RO(STEP)*(ttwo(7,7,1)) increment the time step STEP = STEP + 1 return END *********************** subroutine modelflux(epsil4,t4pos,t4neg) this routine reads the model output and adjusts it based on the heat flux adjustments used in the sensitivity coefficient portion of the program. 125 L _:__zs_ ~ luau-l 0 ()(5 O real(8) tmod(50000),atry(7),epsil4,dumpos,dumneg,flashtime real(8) t4pos(50000),t4neg(50000),zlength integer nread,i common IFLASHIT/ flashtime common IMAINPROG/ tmod,atry common IFOO/ nread common /ZL/ zlength dumpos = atry(4)+epsil4 dumneg = atry(4)-epsil4 re-nondimensionalize the temperature history do i = 1, (3*nread) tmod(i) = tmod(i)/((atry(4)*flashtime*1000)/zlength) t4pos(i) = tmod(i)*((dumpos*flashtime*1000)/zlength) t4neg(i) = tmod(i)*((dumneg*flashtime*1000)/zlength) t4pos(i) = (tmod(i)*dumpos)/atry(4) t4neg(i) = (tmod(i)*dumneg)/atry(4) enddo return end ************************** FUNCTION gasdev(idum) integer idum *uses ran2* returns a normally distributed deviate with zero mean and unit variance, using ran2(idum) as the source of uniform deviates culled from Numerical Recipes in FORTRAN integer iset real(8) fac,gset,rsq,v1,v2 save iset,gset data iset/O/ if(iset.eq.0) then v 1 =2. *ran2(idum)-1. v2=2. *ran2(idum)—1. rsq=v1**2+v2**2 if(rsq.ge. 1 ..or.rsq.eq.0.) goto l fac=sqrt(-2.*log(rsq)/rsq) gset=v1*fac 126 11 gasdev=v2*fac iset=1 else gasdev=gset iset=0 endif return end 316************************* FUNCTION RAN2(IDUM) from Numerical Recipes in FORTRAN returns a uniform random deviate between 0. and 1. PARAMETER (M=714025,IA=1366,IC=150889,RM=1 .4005] 12E-6) DIMENSION IR(97) DATA IFF /O/ IF(IDUM.LT.0.0R.IFF.EQ.O)THEN IFF=1 IDUM=MOD(IC-IDUM,M) DO 1 1 J =1 ,97 IDUM=MOD(IA*IDUM+IC,M) IR(J)=IDUM CONTINUE IDUM=MOD(IA*IDUM+IC,M) IY=IDUM ENDIF J=1+(97*IY)/M IF(J.GT.97.0R.J.LT.1)PAUSE IY=IR(J) RAN2=IY*RM IDUM=MOD(IA*IDUM+IC,M) IR(J)=IDUM RETURN END **************************** end of subprograms, file 127 APPENDIX C RANDOM IN -PLANE AN ISOTROPY FINITE DIFFERENCE PROGRAM 128 The following pages contain FORTRAN code for an explicit finite difference program that models the diffusion equation and assumes monoclinic heat conduction in a slab—like specimen. The thermal diffusivity tensor 0 that allows a monoclinic response may be expressed as 0111 0112 0 a: 0121 0122 0 ((2.1) O O 0.33 where 0.12 = 0121. The parameters to be estimated are the four nonzero components of a. The simulated specimen boundary conditions consist of: insulated edges, convection on the top and bottom surfaces (of different magnitude), and a circular (top surface) heat pulse with a 6 mm radius of constant magnitude and finite duration. The model verifies . that the Courant stability criterion is not violated before each run. The simulated geometry is easily adjustable, with the exception of the geometry of the illuminated area, where the weighting parameters must be adjusted any time the flash diameter is changed. Most of the experimental and geometric parameters are adjusted via the setup subroutine that precedes the actual model program. The output consists of an array containing the temperature history of the simulated specimen at four points on the bottom surface. 129 000000 Sean Davis -- programmer this program computes the temperature history of a single layer specimen that is heated over a portion of the top face. The model handles the fully populated thermal diffusivity tensor, which includes mixed derivatives. The top surface experiences convection, while the other surfaces are either insulated or experience symmetric boundary conditions. subroutine setup(IMAX,stopiter) CHANGE model size here integer NX,NY,NZ parameter (NX = 81) parameter (NY = 81) parameter (NZ = 41) declare variables integer IMAX,sensors,x(5),y(5),z(5),XTC,YTC,nreada,nread,i, j integer dataswitch,noiseswitch,MI,MJ,nn real(8) axx,ayy,azz,axy,pcspec,qest,topconv,botconv,outtime real(8) flashtime,fliz,stopiter,zlength,xlength,ylength,xapt,lxtc real(8) lx,ly,dz,dx,dy,xcorr,ycorr,tdiff(7),timecalc real(8) tdiffo(7),Tmax,noise,qn(NX,NY) character*12 fname common blocks common /GUESS 1/ tdiff,tdiffo common /OTPT/ x,y,z,XTC,YTC,xcorr,ycorr,noise,noiseswitch common IAPT/ qn common IREADITT/ nreada,dataswitch,fliz,nn common /GEOM/ dx,dy,dz,outtime,pcspec common IT [MEI timecalc common /FOO/ nread,lxtc,xapt common /FLASHIT/ flashtime common IFISH/ sensors,Tmax common /FILE/ fname 130 0000000000 000000000 00000 0 set properties -- for inverse problem, these are guesses m1841 parameters azz = 1.1D-07 principal directions (0 degrees) axx = 2.62D-07 ayy = 0.58D-07 axy = 0.0D-07 max shear (45 degrees) axx = 1.6D-07 ayy = 1.6D-07 axy = 1.02D-07 m 1842 parameters azz = 1.1D-07 axx = 1.8D-07 ayy = 1.4D-07 axy = 0.2D-07 axx = 1.883D-07 ayy = 1.317D-07 axy = 0.0D-07 1843 params azz = 1.1D-07 axx = 1.600D-07 ayy = 1.600D-07 axy = 0.564D-07 inverse parameters azz = 1.00D-07 axx = 1.00D-07 ayy = 1.00D-07 axy = 0.50D-07 use axy below if setting shear to zero axy = 0.00D-07 these properties are assumed known pcspec = 1.54D+06 qest = 0.025 topconv = 1.3D-05 botconv = 5./pcspec 131 set experiment parameters fname = "m1843e.txt" timecalc = 100. flashtime = 1.0 fhz = 30. for data input, read every "nn"th point (nn=l means'read every datafile point) nn = 1 outtime = 1./fhz nreada = 3010 IMAX = 50 stopiter = 1.0D-10 sensors = 4 Tmax = 0.0 dataswitch: l = inverse problem, 2 = forward problem dataswitch = 1 forward problem noise parameters -- switch and level noiseswitch: l = noise, 0 = clean data (a noise level of 0.01 = IR Camera) noiseswitch = 0 if (dataswitch.eq.1) noiseswitch = 0 noise = 0.01 set geometry and output information zlength = 3.25 xlength = 40.0 xapt = 6.00 hm=90 zlength = zlength/1000. xlength = xlength/1000. ylength = xlength 1x = lxtc/ 1000. 1y = 1x 132 dz = zlength/dfloat(NZ- 1) dx = xlength/dfloat(NX- l) dy = ylength/dfloat(NY- l) compute center point of model M1 = (NX+1)/2 M1 = (NY+1)/2 for output nodes, "0" nodes are overwritten to account for the inter-nodal sensor locations x(1)= MI x(2) = 0 x(3) = M1 x(4) = 0 x(5) = M1 x(6) = 0 x(7) = 1 y(1)= MI W) = M1 y(3) = 0 y(4) = 0 y(5) = M1 y(6) = 1 y(7) = 0 2(1) = NZ 2(2) = NZ 2(3) = NZ z(4) = NZ 2(5) = 1 2(6) = 1 2(7) = 1 XTC = M1 + int(lx/dx) YTC = M] + int(ly/dy) xcorr = 1. - ((lx - float(XTC-MI)*dx)/dx) ycorr = l. - ((ly - float(YTC-MJ)*dy)/dy) set flux absorption constants -- this models a full circle with a radius of 6.0mm 133 qn=0. do j = MJ-3, MJ+3 do i = MI-l l, MI+11 qn(i,j) = 1. enddo enddo do i = MI-lO, MI+10 qn(i,MJ+4) = 1. qn(i,MJ+5) = 1. qn(i,MJ-4) = 1. qn(i,MJ-S) = 1. enddo do i = MI-9, MI+9 qn(i,MJ+6) = 1. qn(i,MJ+7) = l. qn(i,MJ-6) = 1. qn(i,MJ-7) = l. enddo do i = MI-8, MI+8 qn(i,MJ+8) = 1. qn(i,MJ-8) = 1. enddo do i = MI-7, MI+7 qn(i,MJ+9) = l. qn(i,MJ-9) = 1. enddo do i = MI-5, MI+5 qn(i,MJ+10) = 1. qn(i,MJ-10)= 1. enddo do i = MI-3, MI+3 qn(i,MJ+l 1) = 1. qn(i,MJ-l l) = 1. enddo qn(MI,MJ+12) = 0.5 qn(MI+12,MJ) = 0.5 qn(MI+1 ,MJ+12) = 0.45 134 qn(MI+12,MJ+1) = 0.45 qn(MI+12,MJ+2) = 0.35 qn(MI+2,MJ+12) = 0.35 qn(MI+11,MJ+4) = 0.8 qn(MI+4,MJ+l l) = 0.8 qn(MI+6,MJ+10) = 0.85 qn(MI+10,MJ+6) = 0.85 qn(MI+10,MJ+7) = 0.2 qn(MI+7,MJ +10) = 0.2 qn(MI+9,MJ+8) = 0.5 qn(MI+8,MJ+9) = 0.5 qn(MI+12,MJ+3) = 0.1 qn(MI+3,MJ+12) = 0.1 qn(MI+11,MJ+5) = 0.4 qn(MI+5,MJ+1 1) = 0.4 qn(MI,MJ-12) = 0.5 qn(MI-12,MJ) = 0.5 qn(MI- l ,MJ-12) = 0.45 qn(MI-12,MJ-l) = 0.45 qn(MI-12,MJ-2) = 0.35 qn(MI-2,MJ-12) = 0.35 qn(MI-l 1,MJ-4) = 0.8 qn(MI-4,MJ-1 l) = 0.8 qn(MI—6,MJ -10) = 0.85 qn(MI-10,MJ-6) = 0.85 qn(MI-10,MJ-7) = 0.2 qn(MI-7,MJ-10) = 0.2 qn(MI-9,MJ-8) = 0.5 qn(MI-8,MJ-9) = 0.5 qn(MI-12,MJ-3) = 0.1 qn(MI-3,MJ-l2) = 0.1 qn(MI-l l,MJ-5) = 0.4 qn(MI-5,MJ-l 1) = 0.4 qn(MI+1,MJ-12) = 0.45 qn(MI+12,MJ-l) = 0.45 qn(MI+12,MJ-2) = 0.35 qn(MI+2,MJ-12) = 0.35 qn(MI+1 l,MJ-4) = 0.8 qn(MI+4,M.1-l 1) = 0.8 qn(MI+6,MJ-10) = 0.85 qn(MI+10,MJ-6) = 0.85 qn(MI+10,MJ-7) = 0.2 qn(MI+7,MJ-10) = 0.2 qn(MI+9,MJ-8) = 0.5 135 Onoooono 0 qn(MI+8,MJ-9) = 0.5 qn(MI+12,MJ-3) = 0.1 qn(MI+3,MJ-12) = 0.1 qn(MI+1 l,MJ-5) = 0.4 qn(MI+5,MJ-l 1) = 0.4 qn(MI—1,MJ+12) = 0.45 qn(MI-12,MJ+1) = 0.45 qn(MI-12,MJ+2) = 0.35 qn(MI-2,MJ+12) = 0.35 qn(MI-11,MJ+4) = 0.8 qn(MI-4,MJ+1 l) = 0.8 qn(MI-6,MJ+10) = 0.85 qn(MI-10,MJ+6) = 0.85 qn(MI-10,MJ+7) = 0.2 qn(MI-7,MJ+10) = 0.2 qn(MI-9,MJ+8) = 0.5 qn(MI-8,MJ+9) = 0.5 qn(MI-12,MJ+3) = 0.1 qn(MI-3,MJ+12) = 0.1 qn(MI-11,MJ+5) = 0.4 qn(MI-5,MJ+11) = 0.4 use this to get l-D flash across entire surface doi= 1,NX doj=1,NY qn(i,j)=1 enddo enddo set Marquardt estimation parameters tdiff( 1 )=azz tdiff(2)=axx tdiff(3)=ayy tdiff(4)=axy tdiff(5)=qest tdiff(6)=topconv tdiff(7)=botconv return end 136 0 C ********************************************************************** subroutine model c CHANGE model size here integer NX,NY,NZ parameter (NX = 81) parameter (NY = 81) parameter (NZ = 41) c declare variables c real(8) tmod(25000) real(8) dx,dy,dz,check,qnode,qout,hout,noise,tmod(80000),xapt real(8) T(NX,NY,NZ),Tnew(NX,NY,NZ),Tout(40000,5),timeout(40000) real(8) axx,ayy,azz,axy,pcspec,kzz,delf,bkhout,lxtc,qn(NX,NY) real(8) fxx,fyy,fzz,fxy,hnode,time,outtime,dtf,del,atry(7) real(8) timecalc,flashtime,dt,fliz,xcorr,ycorr,backhnode integer FLASH,FINAL,FINAL2,XTC,YTC,COUNT,nreada,dataswitch,nn integer i,j,k,l,m,x(5),y(5),z(5),noiseswitch,idum,nread,nreadm c common blocks common /MAINPROG/ tmod,atry common /OTPT/ x,y,z,XTC,YTC,xcorr,ycorr,noise,noiseswitch common /APT/ qn common /GEOM/ dx,dy,dz,outtime,pcspec common IT IME/ timecalc common IFLASHIT/ flashtime common IREADI'I'T/ nreada,dataswitch,fhz,nn common IFOO/ nread,lxtc,xapt set the physical parameters for the problem all lengths are in millimeters, all times are in seconds. axx = atry(2) ayy = atry(3) azz = atry( l) axy = atry(4) kzz = azz*pcspec qout = atry(5) 137 _'O-.'h fisl‘p' s .3. ..‘ hout = atry(6) bkhout = atry(7) compute/check the stability criterion for the problem this criteria is the same as that found for the simpler orthogonal problem. check = 1.0 dt = 2./fl1z m = 0 do while (check. ge.0.95) dt = dt/2. fxx = axx*dt/(dx*dx) fyy = ayy*dt/(dy*dy) fzz = azz*dt/(dz*dz) fxy = axy*dt/(2.*dx*dy) check = 2.*fxx + 2.*fyy + 2.*fzz m=m+1 enddo de1= dt/lOO. this section sets the flash timestep dtf = 0.02/fhz check = 1.0 m = 0 do while (check.ge.0.95) dtf = dtf/2. fxx = axx*dtf/(dx*dx) fyy = ayy*dtf/(dy*dy) fzz = azz*dtf/(dz*dz) fxy = axy*dtf/(2.*dx*dy) check = 2.*fxx + 2.*fyy + 2.*fzz m=m+1 enddo delf = dtf/ 1 OO. compute the value of I needed to reach the end of the flash 138 and the end of the problem FLASH = nint(flashtime/dtf) FINAL = nint((timecalc - (2.*flashtime))/dt) FINAL2 = nint(timecalc/dt) compute other dt-dependent parameters qnode = (2.*qout*dtf)/dz hnode = (2.*hout*dtf)/dz backhnode = (2.*bkhout*dtf)/dz COUNT = l T = 0. Tnew = 0. TIME LOOP SHALL BEGIN HERE -- CALCULATION BEGINS loop over experiment time do 1 = 1, 2*FLASH if we are in the flashtime, add the heat flux to the top nodes if(l.le.FLASH) then do i = 1, NX do j = 1, NY T(i,j,l) = T(i,j,1)+qn(i,j)*qnode enddo enddo endif here we compute the temperature field in the body four top comers Tnew(l,l,1) = fxx*(2.*T(2,l,1)-2.*T(l,1,l))+fyy*(2.*T(l,2,1)- + 2.*T(1,1,1))+fzz*(2.*T(],1,2)-2.*T(l,1,1))+T(1,l,1)-hnode*T(l,1,1) Tnew(NX,l,1) = fxx*(2.*T(NX-l,l,1)-2.*T(NX,1,1))+fyy*(2.* + T(NX,2,])-2.*T(NX,1,1))+fzz*(2.*T(NX,l,2)-2.*T(NX,1,1))+T(NX,1,1) 139 + -hnode*T(NX,l , 1) Tnew(l,NY, l) = fxx*(2.*T(2,NY,l)-2.*T(1,NY,1))+fyy*(2.* + T(l,NY-l,l)-2.*T(1 ,NY,1))+fzz*(2.*T(1,NY,2)-2.*T(1,NY,1 ))+ + T(1,NY,1 )-hnode*T(1,NY, 1) Tnew(NX,NY, 1) = fxx*(2.*T(NX-1,NY,])-2.*T(NX,NY,1))+fyy*(2.* + T(NX,NY-1,1)-2.*T(NX,NY,l))+fzz*(2.*T(NX,NY,2)-2.*T(NX,NY,1))+ + T(NX,NY, 1 )-hnode*T(NX,NY, 1) two top edges ‘2 do i = 2, NX-l Tnew(i,1,l) = fxx*(T(i+1,1,1)-2.*T(i,1,1)+T(i-1,1,l))+ + fyy*(2.*T(i,2,1)-2.*T(i,l,l))+fzz*(2.*T(i,l,2)-2.*T(i,l,1))+ 1: + T(i, l , l )-hnode*T(i, l , 1) i1 Tnew(i,NY,1) = fxx*(T(i+1,NY,])-2.*T(i,NY,1)+T(i-l ,NY, 1 ))+ + fyy*(2.*T(i,NY-l ,1)-2.*T(i,NY,1))+fzz*(2.*T(i,NY,2)- + 2.*T(i,NY,1))+T(i,NY,1)-hnode*T(i,NY,1) enddo two top edges doj = 2, NY-l Tnew(l,j,1) = fxx*(2.*T(2,j,1)-2.*T(1,j,1))+ + fyy*(T(l,j+1,l)-2.*T(1,j,1)+T(l,j-1,1))+fzz* + (2.*T(l,j,2)-2.*T(1,j,l))+T(1,j,l)-hnode*T(l,j,l) Tnew(NX,j,1) = fxx*(2.*T(NX-l,j,1)-2.*T(NX,j,1))+ + fyy*(T(NX,j+1,1)-2.*T(NX,j,1)+T(NX,j-l,l))+fzz* + (2.*T(NX,j,2)-2.*T(NX,j,1))+T(NX,j,1)-hnode*T(NX,j,l) enddo top face do i = 2, NX-l do j = 2, NY-l Tnew(i,j,1) = fxx*(T(i+l,j,1)-2.*T(i,j,1)+T(i-1,j,1))+ + fyy*(T(i,j+l,l)-2.*T(i,j,l)+T(i,j-l,1))+fzz*(2.*T(i,j,2)- + 2.*T(i,j,1))+fxy*(T(i+l,j+1,1)—T(i+1,j-1,1)-T(i-1,j+1,1)+ + T(i-l ,j-1,l))+T(i,j,1)-hnode*T(i,j,l) enddo enddo two side faces 140 do i = 2, NX-l do k = 2, NZ-l Tnew(i, 1 ,k) = fxx*(T(i+1,l,k)-2.*T(i,1,k)+T(i-1 ,1,k))+ + fyy*(2.*T(i,2,k)-2.*T(i,1,k))+fzz*(T(i,1,k+l)-2.*T(i,1,k)+ + T(i,] ,k-1))+T(i, 1 ,k) Tnew(i,NY,k) = fxx*(T(i+1,NY,k)-2.*T(i,NY,k)+T(i-1,NY,k))+ + fyy*(2.*T(i,NY-1,k)-2.*T(i,NY,k))+fzz*(T(i,NY,k+1)-2.* + T(i,NY,k)+T(i,NY,k-1))+T(i,NY,k) enddo l enddo 1 two side faces ' do j = 2, NY-l do k = 2, NZ-l Tnew( l ,j,k) = fxx*(2.*T(2,j,k)-2.*T( 1 ,j,k))+fyy* + (T(1,j+1,k)-2.*T(l,j,k)+T(],j-l,k))+fzz*(T(1,j,k+1)-2.* + T(l ,j,k)+T(],j,k-1))+T(l,j,k) Tnew(NX,j,k) = fxx*(2.*T(NX-1,j,k)-2.*T(NX,j,k))+fyy* + (T(NX,j+1,k)-2.*T(NX,j,k)+T(NX,j-l,k))+fzz*(T(NX,j,k+l )- + 2. *T(NX,j,k)+T(NX,j,k- 1 ))+T(NX,j,k) enddo enddo four side edges do k = 2, NZ] Tnew(] , 1,10 = fxx*(2.*T(2,l,k)-2.*T(1,1,k))+fyy* + (2.*T(1,2,k)-2.*T(1,1,k))+fzz*(T(],l,k+1)-2.* + T(1,1,k)+T(1,l,k-1))+T(1,1,k) Tnew(NX, l ,k) = fxx*(2.*T(NX-1,1,k)-2.*T(NX,l,k))+fyy* + (2. *T(NX,2,k)-2.*T(N X, l ,k))+fzz*(T(N X, l ,k+1 )- + 2.*T(NX,1,k)+T(NX,l,k-1))+T(NX,1,k) Tnew( 1 ,NY,k) = fxx*(2. *T(2,NY,k)-2.*T( 1 ,NY,k))-l- + fyy*(2.*T(1,NY-1,k)-2.*T(1,NY,k))+fzz*(T(1,NY,k+1)-2.* + T(l,NY,k)+T(l,NY,k-1))+T(1,NY,k) Tnew(NX,NY,k) = fxx*(2.*T(NX-1,NY,k)-2.*T(NX,NY,k))+ + fyy*(2.*T(NX,NY-1,k)-2.*T(NX,NY,k))+fzz*(T(NX,NY,k+1)-2.* + T(NX,NY,k)+T(NX,NY,k- 1 ))+T(NX,NY,k) enddo 141 model interior do i = 2, NX-l do j = 2, NY-l do k = 2, NZ-l Tnew(i,j,k) = fxx*(T(i+1,j,k)-2.*T(i,j,k)+T(i-1,j,k))+ + fyy*(T(i,j+l ,k)-2.*T(i,j,k)+T(i,j-1 ,k))+fzz* + (T(i,j,k+1)-2.*T(i,j,k)+T(i,j,k-l ))+fxy*(T(i+l ,j+1 ,k)- + T(i+1,j-l,k)-T(i-l,j+1,k)+T(i-1,j-1,k))+T(i,j,k) enddo enddo enddo bottom face do i = 2, NX-l do j = 2, NY-l Tnew(i,j,NZ) = fxx*(T(i+1,j,NZ)-2.*T(i,j,NZ)+T(i-1,j,NZ))+ + fyy*(T(i,j+l ,NZ)-2.*T(i,j,NZ)+T(i,j-l ,NZ))+fzz*(2.* + T(i,j,NZ-l)-2.*T(i,j,NZ))+fxy*(T(i+1,j+1,NZ)- + T(i+1,j-l,NZ)-T(i-l,j+1,NZ)+T(i-1,j-l ,NZ))+T(i,j,NZ)- + backhnode*T(i,j,NZ) enddo enddo two bottom edges do i = 2, NX-l Tnew(i,],NZ) = fxx*(T(i+l,l,NZ)-2.*T(i,],NZ)+T(i-l,1,NZ))+ + fyy*(2.*T(i,2,NZ)-2.*T(i,l ,NZ))+fzz*(2.*T(i,l,NZ-1)- + 2.*T(i, 1 ,NZ))+T(i, 1 ,NZ)-backhnode*T(i, 1 ,NZ) Tnew(i,NY,N Z) = fxx*(T(i+1,NY,NZ)-2.*T(i,NY,NZ)+T(i-1,NY,NZ))+ + fyy*(2.*T(i,NY-l ,NZ)-2.*T(i,NY,NZ))+fzz*(2.*T(i,NY,NZ-1)- + 2.*T(i,NY,NZ))+T(i,NY,NZ)-backhnode*T(i,NY,NZ) enddo two bottom edges do j = 2, NY-l Tnew( l ,j,NZ) = fxx*(2.*T(2,j,NZ)-2.*T( 1 ,j,NZ))+ + fyy*(T(1,j+1,NZ)-2.*T(l,j,NZ)+T(1,j-1,NZ))+fzz* + (2.*T(l,j,NZ-l)-2.*T(1,j,NZ))+T(1,j,NZ)-backhnode*T(1,j,NZ) Tnew(NX,j,NZ) = fxx*(2.*T(NX- l ,j,NZ)-2.*T(NX,j,NZ))+ 142 + + enddo fyy*(T(NX,j+1 ,NZ)-2.*T(NX,j,NZ)+T(NX,j- l ,NZ))+fzz*(2.* T(NX,j,NZ- 1 )-2. *T(NX,j,NZ))+T(N X,j,N Z)-backhnode*T(NX, j,N Z) four bottom comers + + + Tnew( l , l ,N Z) = fxx*(2.*T(2,1,NZ)-2.*T(l ,1,NZ))+fyy*(2.*T(l,2,NZ)- 2.*T(l,1,NZ))+fzz*(2.*T(1,1,NZ-1)-2.*T(1,1,NZ))+T(1,1,NZ)- backhnode*T(1,1 ,NZ) Tnew(NX, 1 ,NZ) = fxx*(2.*T(NX-l,1,NZ)-2.*T(NX,1,NZ))+fyy*(2.* T(NX,2,NZ)-2.*T(NX,l,NZ))+fzz*(2.*T(NX,1,NZ-l)-2.*T(NX,1,NZ))+ T(NX, 1 ,N Z)-backhnode*T(N X, 1 ,NZ) Tnew( 1 ,NY,NZ) = fxx*(2.*T(2,NY,NZ)-2.*T( l ,NY,NZ))+fyy*(2. * T( l ,NY-1 ,NZ)-2.*T(1,NY,NZ))+fzz*(2.*T(1 ,NY,NZ-l )- 2.*T(1,NY,NZ))+T(1,NY,NZ)-backhnode*T(l,NY,NZ) Tnew(NX,NY,NZ) = fxx*(2.*T(NX-1,NY,NZ)- 2. *T(NX,NY,NZ))+fyy*(2. *T(NX,NY- 1 ,NZ)- 2.*T(NX,NY,NZ))+fzz*(2.*T(NX,NY,NZ-1)- 2.*T(NX,NY,NZ))+T(NX,NY,NZ)-backhnode*T(N X,NY,N Z) send Tnew to T once calculations are complete T = Tnew if the calculated time equals the desired output time, write the temperature at the points of interest to a separate array time = dtf*l outtime = (1./fhz)*COUNT if((time+delf).ge.outtime) then timeout(COUNT) = time Tout(COUNT,1) = T(x(1), y(l), z(1)) Tout(COUNT,2) = xcorr*T(XTC,y(2),z(2)) + (1.-xcorr)*T((XTC+1),y(2),z(2)) Tout(COUNT,3) = ycorr*T(x(3),YTC,z(3)) + (1 .-ycorr)*T(x(3),(YTC+1),z(3)) Tout(COUNT,4) = (xcorr*ycorr*T(XTC,YTC,z(4))+(1.-xcorr)* ycorr*T((XTC+1 ),YTC,z(4))+xcorr*( 1 .-ycorr)* 143 00000 + T(XTC,(YTC+1),z(4))+(1.-xcorr)*(1.-ycorr)* + T((XTC+1),(YTC+1),z(4))) Tout(COUNT,5) = T(x(5), y(5), 2(5)) Tout(COUNT,6) = xcorr*T(XTC,y(6),z(6)) + + (1.-xcorr)*T((XTC+l),y(6),z(6)) Tout(COUNT,7) = ycorr*T(x(7),YTC,z(7)) + + (1 .-ycorr)*T(x(7),(YTC+1),z(7)) COUNT = COUNT+1 endif close time loop enddo change parameters to account for larger timestep fxx = axx*dt/(dx*dx) fyy = ayy*dt/(dy*dy) fzz = azz*dt/(dz*dz) fxy = axy*dt/(2.*dx*dy) hnode = (2.*hout*dt)/dz backhnode = (2.*bkhout*dt)/dz loop over rest of time do 1 = 1, FINAL here we compute the temperature field in the body four top comers Tnew(l,1,l) = fxx*(2.*T(2,1,1)-2.*T(1,1,l))+fyy*(2.*T(1,2,l)- + 2.*T(1 ,1,1))+fzz*(2.*T(1,1 ,2)-2.*T(1,1,1))+T(1,1,1)-hnode*T(1,1 ,1 ) Tnew(NX,1,1) = fxx*(2.*T(NX-1,1,1)-2.*T(NX,l,l))+fyy*(2.* + T(NX,2,])-2.*T(NX,1,1))+fzz*(2.*T(NX,l,2)-2.*T(NX,1,1))+T(NX,1,l ) + -hnode*T(NX,1,1) Tnew(l ,NY,1) = fxx*(2.*T(2,NY,l)-2.*T(1,NY,1))+fyy*(2.* + T(1,NY-l,1)-2.*T(1,NY,1))+fzz*(2.*T(1,NY,2)-2.*T(l,NY,1))+ 144 + T( 1 ,N Y, 1 )-hnode*T( 1 ,N Y, 1 ) Tnew(NX,NY,1) = fxx*(2.*T(NX-1 ,NY, l)-2.*T(NX,NY,1 ))+fyy*(2.* T(NX,NY-1,1)-2.*T(NX,NY, l))+fzz*(2.*T(NX,NY,2)-2.*T(NX,NY,1))+ + + T(NX,NY, 1 )-hnode*T(NX,NY, 1) two top edges do i = 2, NX-l Tnew(i,1,1)= fxx*(T(i+1,1,1)-2.*T(i,l ,1)+T(i-1,1,1))+ + fyy*(2.*T(i,2,])-2.*T(i,1,1))+fzz*(2.*T(i,1,2)-2.*T(i,l,l))+ + T(i, 1 ,1)-hnode*T(i,1,1) Tnew(i,NY, l) = fxx*(T(i+1,NY,l)-2.*T(i,NY,l)+T(i-1,NY,1))+ + fyy*(2.*T(i,NY-1,1)-2.*T(i,NY,1))+fzz*(2.*T(i,NY,2)- + 2.*T(i,N Y, 1 ))+T(i,N Y, 1 )-hnode*T(i,N Y, 1 ) enddo two top edges do j = 2, NY-l Tnew(l,j,1) = fxx*(2.*T(2,j,1)-2.*T(1,j,l))+ + fyy*(T(l,j+1,l)-2.*T(1,j,l)+T(1,j-l,1))+fzz* + (2.*T(l,j,2)—2.*T(l,j,1))+T(1,j,l)-hnode*T(1,j,1) Tnew(NX,j, 1) = fxx*(2.*T(NX- l ,j,l )-2.*T(NX,j, 1))+ + fyy*(T(NX,j+l ,1)-2.*T(NX,j,1)+T(NX,j-l,1))+fzz* + (2.*T(NX,j,2)-2.*T(NX,j,1))+T(NX,j,1)-hnode*T(NX,j,1) enddo top face do i = 2, NX-l do j = 2, NY-l Tnew(i,j,l) = fxx*(T(i+1,j,1)-2.*T(i,j,1)+T(i-l,j,1))+ + fyy*(T(i,j+1,1)-2.*T(i,j,l)+T(i,j-l,1))+fzz*(2.*T(i,j,2)- + 2.*T(i,j,1))+fxy*(T(i+1,j+l,l)-T(i+1,j-1,1)-T(i-1,j+l,l)+ + T(i-l,j-l,1))+T(i,j,l)-hnode*T(i,j,l) enddo enddo two side faces doi=2,NX-1 do k = 2, NZ-l 145 Tnew(i, 1 ,k) = fxx*(T(i+1,1,k)-2.*T(i,1,k)+T(i-1,l,k))+ + fyy*(2.*T(i,2,k)-2.*T(i,l,k))+fzz*(T(i,l,k+1)-2.*T(i,1,k)+ + T(i,],k-1))+T(i,1,k) Tnew(i,NY,k) = fxx*(T(i+1,NY,k)-2.*T(i,NY,k)+T(i-l,NY,k))+ + fyy*(2.*T(i,NY-1 ,k)-2.*T(i,NY,k))+fzz*(T(i,NY,k+1)-2.* + T(i,NY,k)+T(i,NY,k-1))+T(i,NY,k) enddo enddo two side faces do j = 2, NY-l do k = 2, NZ-l Tnew( 1 ,j,k) = fxx*(2.*T(2,j,k)-2.*T(1 ,j,k))+fyy* + (T(l,j+l,k)-2.*T(l,j,k)+T(],j-l,k))+fzz*('1‘(l,j,k+1 )-2.* + T(1,j,k)+T(l,j,k-1))+T(1,j,k) Tnew(NX,j,k) = fxx*(2.*T(NX- l ,j,k)-2.*T(NX,j,k))+fyy* + (T(NX,j+l,k)-2.*T(NX,j,k)+T(NX,j-l,k))+fzz*(T(NX,j,k+1)- + 2.*T(NX,j,k)+T(NX,j,k-l))+T(NX,j,k) enddo enddo four side edges do k = 2, NZ] Tnew(1,1,k) = fxx*(2.*T(2,1,k)-2.*T(1,l,k))+fyy* + (2.*T(1,2,k)-2.*T(l,1,k))+fzz*(T(1,1,k+l)-2.* + T(l,1,k)+T(1,1,k-l))+T(l,1,k) Tnew(N X, 1 ,k) = fxx*(2.*T(NX-1,1,k)-2.*T(NX,1 ,k))+fyy* + (2.*T(NX,2,k)-2.*T(NX,1,k))+fzz*(T(NX,1,k+1)- + 2.*T(NX,1,k)+T(NX,1,k-l))+T(NX,1,k) Tnew(l,NY,k) = fxx*(2.*T(2,NY,k)-2.*T(l ,NY,k))+ + fyy*(2.*T(1,NY-1,k)-2.*T(1,NY,k))+fzz*(T(1,NY,k+1)-2.* + T(l,NY,k)+T(l,NY,k-1))+T(1,NY,k) Tnew(NX,NY,k) = fxx*(2.*T(NX-1,NY,k)-2.*T(NX,NY,k))+ + fyy*(2.*T(NX,NY-l,k)-2.*T(NX,NY,k))+fzz*(T(NX,NY,k+1)-2.* + T(NX,NY,k)+T(NX,NY,k-l))+T(NX,NY,k) enddo model interior 146 do i = 2, NX-l do j = 2, NY-l do k = 2, NZ-l Tnew(i,j,k) = fxx*(T(i+1,j,k)-2.*T(i,j,k)+T(i-l,j,k))+ + fyy*(T(i,j+ 1 ,k)-2.*T(i,j,k)+T(i,j- l ,k))+fzz* + (T(i,j,k+1)-2.*T(i,j,k)+T(i,j,k-1))+fxy*(T(i+1,j+l ,k)- + T(i+1,j-1,k)-T(i—1,j+l,k)+T(i-1,j-1,k))+T(i,j,k) enddo enddo enddo bottom face do i = 2, NX-l do j = 2, NY-l Tnew(i,j,NZ) = fxx*(T(i+1,j,NZ)-2.*T(i,j,NZ)+T(i-l,j,NZ))+ + fyy*(T(i,j+],NZ)-2.*T(i,j,NZ)+T(i,j-1,NZ))+fzz*(2.* + T(i,j,NZ-1 )-2.*T(i,j,NZ))+fxy*(T(i+l ,j+1 ,NZ)- + T(i+1,j-l ,NZ)-T(i-1,j+l,NZ)+T(i-l,j-1,NZ))+T(i,j,NZ)- + backhnode*T(i,j,NZ) enddo enddo two bottom edges do i = 2, NX-l Tnew(i,1,NZ) = fxx*(T(i+1,1,NZ)-2.*T(i,],NZ)+T(i-1,1,NZ))+ + fyy*(2.*T(i,2,NZ)-2.*T(i,l ,NZ))+fzz*(2.*T(i,1,NZ-1)- + 2.*T(i,],NZ))+T(i,1,NZ)-backhnode*T(i,l,NZ) Tnew(i,NY,N Z) = fxx*(T(i+1,NY,NZ)-2.*T(i,NY,NZ)+T(i-l,NY,NZ))+ + fyy*(2.*T(i,NY-1,NZ)-2.*T(i,NY,NZ))+fzz*(2.*T(i,NY,NZ-1)- + 2.*T(i,NY,NZ))+T(i,NY,NZ)-backhnode*T(i,NY,NZ) enddo two bottom edges do j=2, NY-l Tnew(l,j,NZ) = fxx*(2.*T(2,j,NZ)-2.*T(l,j,NZ))+ + fyy*(T(l,j+1,NZ)-2.*T(1,j,NZ)+T(],j-1,NZ))+fzz* + (2.*T(],j,NZ-l)-2.*T(1,j,NZ))+T(l,j,NZ)-backhnode*T(l,j,NZ) Tnew(NX,j,NZ) = fxx*(2.*T(NX-1,j,NZ)-2.*T(NX,j,NZ))+ + fyy*(T(NX,j-I-1,NZ)-2.*T(NX,j,NZ)+T(NX,j-l ,NZ))+fzz*(2.* + T(NX,j,N Z- 1 )-2.*T(NX,j,NZ))+T(NX,j,NZ)—backhnode*T(NX,j ,N Z) 147 enddo four bottom comers + + + Tnew(1,l ,NZ) = fxx*(2.*T(2,l,NZ)-2.*T(l,1,NZ))+fyy*(2.*T(],2,NZ)— 2.*T(l,l,NZ))+fzz*(2.*T(1,1,NZ—1)-2.*T(1,1,NZ))+T(1,1,NZ)— backhnode*T( l ,1 ,NZ) Tnew(NX, 1 ,NZ) = fxx*(2.*T(NX-1,1,NZ)-2.*T(NX,1,NZ))+fyy*(2.* T(NX,2,NZ)—2.*T(NX,1,NZ))+fzz*(2.*T(NX,1,NZ-l)-2.*T(NX,1,NZ))+ T(NX,1,NZ)-backhnode*T(NX,1 ,NZ) Tnew( 1 ,NY,NZ) = fxx*(2.*T(2,NY,NZ)-2.*T(1,NY,NZ))+fyy*(2.* T(l,NY-1,NZ)-2.*T(l,NY,NZ))+fzz*(2.*T(1,NY,NZ-1)- 2.*T(1,NY,NZ))+T(1,NY,NZ)-backhnode*T(1,NY,NZ) Tnew(NX,NY,NZ) = fxx*(2.*T(NX-1,NY,NZ)- 2.*T(NX,NY,NZ))+fyy*(2.*T(NX,NY-1,NZ)- 2.*T(NX,NY,NZ))+fzz*(2.*T(NX,NY,NZ-1)- 2.*T(NX,NY,NZ))+T(NX,NY,NZ)-backhnode*T(NX,NY,NZ) send Tnew to T once calculations are complete T = Tnew if the calculated time equals the desired output time, write the temperature at the points of interest to a separate array time = dt*l + 2.*flashtime outtime = ( l .lfliz)*COUNT if((time+del).ge.outtime) then + timeout(COUNT) = time Tout(COUNT,1) = T(x(1), y(l), 2(1)) Tout(COUNT,2) = xcorr*T(XTC,y(2),z(2)) + (1 .-xcorr)*T((XTC+l),y(2),z(2)) Tout(COUNT,3) = ycorr*T(x(3),YTC,z(3)) + (1 .-ycorr)*T(x(3),(YTC+l ),z(3)) Tout(COUNT,4) = (xcorr*ycorr*T(XTC,YTC,z(4))+(1 .-xcorr)* ycorr*T((XTC+1),YTC,z(4))+xcorr*(l .-ycorr)* T(XTC,(YTC+1),z(4))+(1.-xcorr)*(l.-ycorr)* T((XTC+1),(YTC+1),z(4))) 148 00000 0 Tout(COUNT,5) = T(x(5), y(5), 2(5)) Tout(COUNT,6) = xcorr*T(XTC,y(6),z(6)) + + (1 .-xcorr)*T((XTC+l),y(6),z(6)) Tout(COUNT,7) = ycorr*T(x(7),YTC,z(7)) + + (1 .-ycorr)*T(x(7),(YTC+1),z(7)) outtime = outtime + (1 ./f1‘lZ) COUNT = COUNT+1 endif close time loop enddo nreadm = COUNT-l handle program output send output temperature history to the MO program do i = 1, nreadm j = i+nreadm k = i+2*nreadm l = i+3*nreadm set to center TC tmod(i) = Tout(i,1) set to old X TC tmod(i) = Tout(i,2) set to old Y TC tmod(k) = Tout(i,3) set to off axis sensor tmod(l) = Tout(i,4) enddo if we are running the forward problem, add noise to simulate experimental data. if (noiseswitch.eq. 1) then call SYSTEM_CLOCK(idum) if (idum.eq.0) pause 'idum = 0; restart' write(*,*) 'idum = ',idum do i = 1,nreadm do j = 1,5 149 0 ()(5 0 95 Tout(i,j) = Tout(i,j) + noise*gasdev(idum) enddo enddo elseif(noiseswitch.eq.0) then confinue else pause 'error with noiseswitch' endif open output file and write out temperature history at the points of interest open (unit=92, file="modelout.txt", status = "unknown") do 1 = 1, nreadm write(92,95) timeout(l),Tout(1, 1 ),Tout(l,2),Tout(l,3), + Tout(l,4),Tout(l,5) enddo format (1X,F7.3,2X,5F9.3) close files, return to main program close(92) return end ************************** FUNCTION gasdev(idum) integer idum *uses ran2* returns a normally distributed deviate with zero mean and unit variance, using ran2(idum) as the source of uniform deviates culled from Numerical Recipes in FORTRAN integer iset real(8) fac,gset,rsq,v 1 ,v2 save iset,gset data iset/OI if(iset.eq.0) then v 1 =2.*ran2(idum)-1. v2=2.*ran2(idum)-1. rsq=v1**2+v2**2 if(rsq.ge. l ..or.rsq.eq.0.) goto 1 150 11 0 fac=sqrt(-2.*log(rsq)/rsq) gset=vl *fac gasdev=v2*fac iset: 1 else gasdev=gset iset=0 endif return end ************************** FUNCTION RAN2(IDUM) from Numerical Recipes in FORTRAN returns a uniform random deviate between 0. and l. PARAMETER (M=7 14025,IA=1 366,IC=150889,RM=1 .4005 1 12E-6) DIMENSION IR(97) DATA IFF /0/ IF(IDUM.LT.0.0R.IFF.EQ.0)THEN IFF=1 IDUM=MOD(IC-IDUM,M) DO 11 J=l,97 IDUM=MOD(IA*IDUM+IC,M) IR(J)=IDUM CONTINUE IDUM=MOD(IA*IDUM+IC,M) IY=IDUM ENDIF J=1+(97*IY)/M IF(J.GT.97.0R.J.LT.1)PAUSE IY=IR(J) RAN2=IY*RM IDUM=MOD(IA*IDUM+IC,M) IR(J)=IDUM RETURN END **************************** end of subprograms, file 151 REFERENCES 152 REFERENCES Aviles-Ramos, C., 2002, “Exact Solution of Heat Conduction in a Two-Domain Parallelepiped with an Orthotropic Layer and its Application to the Estimation of the Three-Dimensional Thermal Conductivity Tensor and Volumetric Heat Capacity of the Orthotropic Layer,” Proceedings of the ASME IMECE 2002, New Orleans, Louisiana, November 17-22 2002, IMECE2002-32424. Beck J .V., and Arnold, K.J., 1977, Parameter Estimation in Engineering and Science, John Wiley and Sons, New York. Bennett, T.D., 2004, “Determining Anisotropic Film Thermal Properties Through Harmonic Surface Heating With a Gaussian Laser Beam: A Theoretical Consideration,” Journal of Heat Transfer 126, pp. 305-31 1. Bradley, CB, 1943, “Measurement of thermal conductivity of thermal insulating materials by means of guarded hot plate,” Refrigerating Engineering 45(5), pp. 337-341. Broerman, A.W., Venerus, DC, and Schieber, J.D., 1999, “Evidence for the Stress- Therrnal Rule in an Elastomer Subjected to Simple Elongation,” Journal of Chemical Physics, 111(15), pp. 6965-6969. Carslaw, HS, and J aeger, J.C., 1959, Conduction of Heat in Solids, 2"d ed., Oxford University Press, Oxford, UK. Chang, Y.P., and Tsou, R.C.H., 1977, “Heat Conduction in an Anisotropic Medium Homogeneous in Cylindrical Regions - Unsteady State,” Journal of Heat Transfer, 99, pp. 41-46. Choy, C.L., Luk, W.H., and Chen, EC, 1978, “Thermal Conductivity of Highly Oriented Polyethylene,” Polymer, 19(2), pp. 155-162. Chu, F.I., Taylor, RE, and Donaldson, A.B., 1980, “Thermal Diffusivity Measurements at High Temperatures by the Radial Flash Method,” Journal of Applied Physics, 51(1) pp. 336-341. CRC Handbook of Chemistry and Physics, 71St ed., D. Lide, editor, CRC Press, Boston. Davis, SE, and Wright, N.T., 2005, “An Optimal Experimental Design Approach for Sensor Positioning in an Extended Flash Thermal Diffusivity Experiment”, in preparation. 153 Degiovanni, A., 1977, “Diffusivite et methode flash (Thermal Diffusivity and the Flash Method),” Revue Generale de Thermique, 16(185), pp. 420—441. Donaldson, A.B., and Taylor, R.B., 1975, “Thermal Diffusivity Measurement by a Radial Heat Flow Method,” Journal of Applied Physics, 46 (10), pp. 4584-4589. Doss, D.J., 1998, “Thermal Diffusivity of Vascular Tissue and Polymers: Effects of Finite Deformation and Direction,” Ph.D. dissertation, University of Maryland Baltimore County, Baltimore, MD. Doss, D.J., and Wright, N.T., 2000, “Simultaneous Measurement of the Orthogonal Components of Thermal Diffusivity in PVC Sheet,” Journal of Heat Transfer, 122(1), pp. 27-32. Douglas, J ., 1962, “Alternating Direction Methods for Three Space Variables,” Numerische Mathematik, 41, pp. 41-63. Dowding, K.J., and Blackwell, BE, 1999, “Design of Experiments to Estimate Temperature Dependent Thermal Properties,” Proceedings of ASME 3” Conference on Inverse Problems in Engineering: Theory and Practice, pp. 509- 5 18. Emery, A.F., 2001, “Using the Concept of Information to Optimally Design Experiments with Uncertain Parameters,” Journal of Heat Transfer, 123, pp. 593-600. Emery, A.F., and Fadale, T.D., 1997, “The Effect of Imprecisions in Thermal Sensor Location and Boundary Conditions on Optimal Sensor Location and Experimental Accuracy,” Journal of Heat Transfer, 119, pp. 661-665. Emery, A.F., Nenarokomov, A.V., and Fadale, T.D., 2000, “Uncertainties in Parameter Estimation: The Optimal Experiment Design,” International Journal of Heat and Mass Transfer, 43, pp. 3331-3339. Fadale, T.D., Nenarokomov, A.V., and Emery, A.F., 1995, “Two Approaches to Optimal Sensor Locations,” Journal of Heat Transfer, 117, pp. 373-379. Graham, 8., McDowell, D.L., and Dinwiddie, R.B., 1999, “In-Plane Thermal Diffusivity Measurements of Orthotropic Materials,” Thermal Conductivity 24, RS. Gaal, ed., Technomic, Lancaster, PA, pp. 241-252. Graham, 8., McDowell, D.L., and Dinwiddie, R.B., 1999, “Multidimensional Flash Diffusivity Measurements of Orthotropic Materials,” International Journal of Thermophysics, 20(2), pp. 691-707. 154 Haji-Sheikh, A., and Beck, J.V., 2002, “Temperature Solution in Multi-Dimensional Multi-Layer Bodies,” International Journal of Heat and Mass Transfer, 45, pp. 1865- 1877. Hsieh, M., and Ma, Q, 2002, “Analytical Investigations for Heat Conduction Problems in Anisotropic Thin-Layer Media With Embedded Heat Sources,” International Journal of Heat and Mass Transfer, 45, pp. 4117-4132. Huang, SC, and Chang, Y.P., 1984, “Anisotropic Heat Conduction With Mixed Boundary Conditions,” Journal of Heat Transfer, 106, pp. 646-648. Huang, X., 2004, “Comparison of Thermal Diffusivity Measurement Techniques,” M.S. thesis, Michigan State University, East Lansing, MI. Incropera, F., and DeWitt, D., 1990, Fundamentals of Heat Transfer, 3'“ edition, John Wiley and Sons, New York. Kato, H, Baba, T, and Okaji, M, 2001, “Anisotropic Thennal-Diffusivity Measurements by a New Laser-Spot-Heating Technique,” Measurement Science and Technology, 12, pp. 2074-2080. LeGall, E.L., 2001, “Thermal Diffusivity of Elastomers Subject to Finite Biaxial Deformation,” Ph.D. dissertation, University of Maryland Baltimore County, Baltimore, MD. LeGall, E.L., and Wright, N.T., 2000, “Changes in Thermal Diffusivity of Rubber Due to Mechanical Preconditioning”, paper presented at the Fourteenth Symposium on Thermophysical Properties, June 2000, Boulder Colorado, pp. 1-22. Le Niliot, C., Rigollet, F., and Petit, D., 2000, “An Experimental Identification of Line Heat Sources in a Diffusive System Using the Boundary Element Method,” International Journal of Heat and Mass Transfer, 43, pp. 2205-2220. Ma, C. and Chang, S., 2004, “Analytical Exact Solutions of Heat Conduction Problems for Anisotropic Multi-Layered Media,” International Journal of Heat and Mass Transfer, 47, pp. 1643-1655. Maillet, D., Lachi, M. and Degiovanni, A., 1990, “Simultaneous Measurements of Axial and Radial Thermal Diffusivities of an Anisotropic Solid in Thin Plate: Application to Multi-Layered Materials,” Thermal Conductivity 21, Cremers, OJ. and Fine, H.A., eds., Plenum Press, New York, pp. 91-107. Mandelis, A., 1985, “Hamilton-Jacobi Formulation and Quantum Theory of Thermal Wave Propagation in the Solid State,” Journal of Mathematical Physics, 26(10), pp. 2676-2683. 155 Mandelis, A., Funak, F., and Munidasa, M., 1996, “Generalized Methodology for Thermal Diffusivity Depth Profile Reconstruction in Semi-infinite and Finitely Thick Inhomogeneous Solids,” Journal of Applied Physics, 80(10), pp. 5570- 5578. Matthews, G., 1996, PVC: Production, Properties and Uses, The University Press, Cambridge. McMasters, R.L., 1997, “Modeling Heat Transfer for Parameter Estimation in Flash Diffusivity Experiments,” Ph.D. dissertation, Michigan State University, East Lansing, MI. McMasters, R.L., Beck, J.V., Dinwiddie, R.B., and Wang, H., 1999, “Accounting for Penetration of Laser Heating in Flash Thermal Diffusivity Experiments,” Journal of Heat Transfer, 121, pp. 15-21. Mejias, M.M., Orlande, H.R.B., and Ozisik, M.N., 1999, “Design of Optimum Experiments for the Estimation of the Thermal Conductivity Components of Orthotropic Solids,” Hybrid Methods in Engineering, 1, pp. 37-53. Mejias, M.M., Orlande, H.R.B., and Ozisik, M.N., 2000, “On the Choice of Boundary Conditions for the Estimation of the Thermal Conductivity Components of Orthotropic Solids,” Proceedings of NH TC ’00, 34m National Heat Transfer Conference, Pittsburgh, Pennsylvania, August 20-22, 2000, NHTC2000-12062. Mejias, M.M., Orlande, H.R.B., and Ozisik, M.N., 2003, “Effects of the Heating Process and Body Dimensions on the Estimation of the Thermal Conductivity Components of Orthotropic Solids,” Inverse Problems in Engineering, 11(1), pp. 75-89. Milosevic, ND, and Raynaud, M., 2004, “Analytical Solution of Transient Heat Conduction in a Two-Layer Anisotropic Cylindrical Slab Excited Superficially by a Short Laser Pulse,” International Journal of Heat and Mass Transfer, 47 , pp. 1627-1641. Mulholland, GP, and Gupta, BR, 1977, “Heat Transfer in a Three-Dimensional Anisotropic Solid of Arbitrary Shape,” Journal of Heat Transfer, 99, pp. 135-137. Mzali, F., Sassi, L., Jemni, A., Ben Nasrallah, S., and Petit, D., 2004, “Optimal Experiment Design for the Identification of Thermo-Physical Properties of Orthotropic Solids,” Inverse Problems in Science and Engineering, 12(2), pp. 193-209. Onsager, L., 1931, “Reciprocal Relations in Irreversible Processes. 1,” Physical Review, 37(4) pp. 405-426. 156 Ozisik, M.N., 1993, Heat Conduction, 2'“! ed., John Wiley and Sons, Inc., New York. Parker, W.J., Jenkins, R.J., Butler, CR, and Abbott, G.L., 1961, “Flash Method of Determining Thermal Diffusivity, Heat Capacity, and Thermal Conductivity,” Journal of Applied Physics, 32(9), pp. 1679-1684. Petit, D., and Hachette, R., 1998, “Model Reduction in Linear Heat Conduction: Use of Interface Fluxes for the Numerical Coupling,” International Journal of Heat and Mass Transfer, 41, pp. 3177-3189. Philippi, 1., Batsale J .C., Maillet, D., and Degiovanni, A., 1995, “Measurement of Thermal Diffusivities Through Processing of Infrared Images,” Review of Scientific Instruments, 66(1), pp. 182-192. Poon, K.C., Tsou, R.C.H., and Chang, Y.P., 1979, “Solution of Anisotropic Problems of First-Class by Coordinate Transformation,” Journal of Heat Transfer, 101, pp. 340-345. Powers, J.M., 2004, “On the Necessity of Positive Semi-Definite Conductivity and Onsager Reciprocity in Modeling Heat Conduction in Anisotropic Media,” Journal of Heat Transfer, 126, pp. 670-675. Pozrikidis, C., 1998, Numerical Computation in Science and Engineering, New York: Oxford University Press, pp. 526-550. Press, W.H., Teukolsky, S.A., Vetterling, WT, and Flannery, BR, 1992, Numerical Recipes in FORTRAN: The Art of Scientific Computing, 2"d ed., Cambridge University Press, New York. Pron, H., and Bissieux, C., 2004, “3-D Thermal Modelling Applied to Stress-Induced Anisotropy of Thermal Conductivity,” International Journal of Thermal Sciences, 43, pp. 1161-1169. Sawaf, B., and Ozisik, M.N., 1995, “Determining the Constant Thermal Conductivities of Orthotropic Materials by Inverse Analysis,” International Communications on Heat and Mass Transfer, 22(2), pp. 201-211. Taktak, R., 1992, “Design and Validation of Optimal Experiments for Estimating Thermal Properties of Composite Materials,” Ph.D. dissertation, Michigan State University, East Lansing, MI. Taktak, R., Beck, J.V., and Scott, ER, 1993, “Optimal Experimental Design for Estimating Thermal Properties of Composite Materials,” International Journal of Heat and Mass Transfer, 36(12), pp. 2977-2986. 157 Torres, J.H., Springer, T.A., Welch, A.J., and Pearce, J .A., 1990, “Limitations of a Thermal Camera in Measuring Surface Temperature of Laser Irradiated Tissues,” Lasers in Surgery and Medicine, 10, pp. 510-523. Venerus, D.C., Schieber, J.D., Iddir, H., Guzman, J .D., and Broerman, A.W., 1999, “Relaxation of Anisotropic Thermal Diffusivity in a Polymer Melt Following Step Shear Strain,” Physical Review Letters, 82(2), pp. 366-369. Venerus, D.C., Schieber, J .D., Iddir, H., Guzman, J.D., and Broerman, A.W., 1999, “Measurement of Thermal Diffusivity in Polymer Melts Using Forced Rayleigh Light Scattering,” Journal of Polymer Science Part B: Polymer Physics, 37, pp. 1069- 1078. Venerus, D.C., Schieber, J.D., Iddir, H., Guzman, J ., and Broerman, A., 2001, “Anisotropic Thermal Diffusivity Measurements in Defonning Polymers and the Stress-Thermal Rule,” International Journal of Thermophysics, 22(4) pp. 1215- 1225. Videcoq, E., and Petit, D., 2001, “Model Reduction for the Resolution of Multidimensional Inverse Heat Conduction Problems,” International Journal of Heat and Mass Transfer, 44, pp. 1899-1911. 158