Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/245363171

Numerical Determination of Radiative View Factors Using Ray Tracing

Article in Journal of Heat Transfer · July 2010


DOI: 10.1115/1.4000974

CITATIONS READS

16 428

3 authors, including:

Shicheng Xue
The University of Sydney
41 PUBLICATIONS 649 CITATIONS

SEE PROFILE

All content following this page was uploaded by Shicheng Xue on 02 August 2015.

The user has requested enhancement of the downloaded file.


Numerical Determination of
Radiative View Factors Using Ray
T. Walker
Tracing
S.-C. Xue
A ray-tracing method is presented for numerically determining radiative view factors in
G. W. Barton complex three-dimensional geometries. This method uses a set of “primitive” shapes to
e-mail: gbarton@usyd.edu.au approximate the required geometry together with a Monte Carlo simulation to track the
fate of randomized rays leaving each surface. View factors were calculated for an op-
School of Chemical and Biomolecular erational fiber drawing furnace using both numerical integration and ray-tracing meth-
Engineering, ods. Calculated view factor profiles were essentially identical above a ray density of 105
University of Sydney, per unit area. Run times for the ray-tracing method were considerable longer, although
New South Wales 2006, Australia the setup time to describe a new geometry is very short and essentially independent of
system complexity. 关DOI: 10.1115/1.4000974兴

Keywords: radiative heat transfer, view factor, ray tracing, Monte Carlo method, poly-
mer optical fiber

1 Introduction 共i.e., nondeforming兲 preform and used in a heat transfer model of


the drawing furnace with close agreement being obtained between
Figure 1 shows an arbitrarily orientated surface Ai having a
simulated and experimental 共measured at the preform surface and
constant total hemispherical emissivity ␧i and temperature Ti. Of center兲 temperature profiles. These view factors are presented later
the total radiant energy leaving this surface in all directions to confirm the accuracy of results obtained using an alternative
共␧i␴0T4i Ai, where ␴0 is the Stefan–Boltzmann constant兲, only a ray-tracing method.
fraction will reach a second arbitrarily orientated surface A j. This The remainder of this paper is thus structured as follows. Sec-
fraction 共Fi-j兲 is defined as the view factor, a term that is consis- tion 2 presents the concept of ray-tracing and how “primitives”
tent with the fact that its value is dependent only on the geometri- can be used to calculate view factors. The tapered cylinder primi-
cal orientation of the two participating surfaces. A view factor can tive is presented in detail, as this is used in calculating view fac-
thus be expressed 共and calculated兲 as follows: tors for the case where a cylindrical preform is drawn to fiber.

冕 冉冕 冊
Section 3 presents view factor results for both the nondeforming
1 cos ␪i cos ␪ j preform and the tapering preform cases. In the former case, both
Fi-j = dA j dAi 共1兲
Ai ␲Si-j
2 numerical integration and ray tracing are used, while in the latter
only ray tracing is employed. This section also compares the rela-
where dAi and dA j are differential elements within the two sur- tive merits of the two methods.
faces with Si-j being the length of the connecting line between the
two, while ␪i and ␪ j are the respective angles between the con-
necting line and the normal vectors 共ni and n j兲 to each surface.
Evaluation of view factors is straightforward provided that Si-j,
cos ␪i, and cos ␪ j can be expressed in terms of the geometrical 2 Ray Tracing
parameters that define the two participating surfaces and the nec- Ray tracing is a widely used technique in optics whereby the
essary integration performed. For radiative exchange between an path of a photon or “ray” is followed through a three-dimensional
infinitesimal and a finite surface, only a single area integration is 共3D兲 environment from its point of origin to its final destination.
needed. However, if two finite surfaces are involved, then a This technique has been employed to solve diverse problems from
double area integral is required. Because of the importance of the generation of images with a high degree of photorealism 关3,4兴
radiative heat transfer in a wide variety of applications, compila- to complex radiative heat transfer modeling 关5–10兴. In the present
tions of analytical or tabulated results 共often in terms of dimen- context, ray tracing may be adapted to calculate the view factor
sionless geometrical parameters兲 are available in the literature 关1兴. between objects through its implementation within a Monte Carlo
In some cases, an unknown view factor can be generated from simulation 关11兴. Essentially a large number of rays are “fired”
known factors by making use of view factor algebra 关2兴. from random points on a given object at random angles into the
The present work was motivated by the need to calculate the 3D environment with the first object that each ray intersects being
radiative heat transfer within an operational furnace used to heat a recorded. The view factor between any two objects may be de-
polymer preform sufficiently to be drawn to fiber. Here the avail- fined as the fraction of rays leaving one object that reaches the
able view factor compilations were of limited use due to the geo- other.
metric complexity of our drawing furnace, and the fact that once Accurate view factor estimation here requires that careful atten-
the glass transition temperature is reached then the shape of the tion be paid to a number of computational issues, the two most
deforming preform becomes a strong function of the thermal con- important being the distribution of 共random兲 points on the emit-
ditions within the furnace. Initially all view factors were deter- ting surface and their 共random兲 emission into the 3D environment.
mined using numerical integration for the case of a cylindrical The major strength of ray tracing as a tool for numerically
estimating view factors is that its application is essentially inde-
Contributed by the Heat Transfer Division of ASME for publication in the JOUR-
pendent of the complexity of the 3D geometry that defines the
HEAT TRANSFER. Manuscript received May 31, 2009; final manuscript received
NAL OF system. By comparison, alternatives such as numerical integration
November 26, 2009; published online April 28, 2010. Assoc. Editor: He-Ping Tan. become increasingly challenging with geometrical complexity.

Journal of Heat Transfer Copyright © 2010 by ASME JULY 2010, Vol. 132 / 072702-1

Downloaded 29 Apr 2010 to 129.78.64.102. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Fig. 2 Procedure for transforming a ray to find ray-object
intersection
Fig. 1 Configuration for radiative heat exchange between two
finite surfaces

can be described in terms of a stack of tapered cylinders. The


In passing, it should be noted that although the focus of this description of this primitive is thus considered next.
paper is on the use of ray tracing in view factor calculation, the
technique can be conceptually expanded to calculate net radiative 2.2 Tapered Cylinder Primitive. The primitive here is a cyl-
heat transfer within any complex 3D enclosure 关12兴. Here each inder with a base radius of 1 and a variable upper radius. As such,
ray emitted from a surface is taken as representing a “packet” of it can be used to model a wide range of shapes from a cylinder,
thermal energy that is either absorbed 共and possibly re-emitted兲 or through a conical frustum, to a cone. The “surface equation” for
reflected when impinging on a second surface. However, it must this primitive when its base is centered on the origin in the x-y
be said that our 共unpublished兲 case studies have shown this to be plane and it extends along the z axis to a height of 1 is as follows:
a computationally demanding approach. x2 + y 2 − 共1 + 共s − 1兲z兲2 = 0 共2兲
2.1 Use of Primitives in Ray Tracing. Ray tracing is not a where s is the radius of the upper surface such that 0 ⱕ s ⱕ 1.
new approach for calculating view factors with a literature dating When ray tracing is used to calculate view factors, it is essential
back some 40 years 关13,14兴. However, most approaches advocate to ensure an even distribution of ray starting points from the sur-
that any object within a system be described via a mesh which face. For a simple cylinder 共for which s = 1兲, ray starting points
maps to its surface 关11,15,16兴. While this approach reduces the can be defined in terms of the azimuthal angle 共␾兲 as follows:
complexity of calculating suitable ray launch positions and launch
directions, it can introduce representational errors in terms of ad- z=␹ 共3兲
equately modeling an object’s curvature and requires a robust dis-
cretization algorithm. x = rz cos共␾兲 共4兲
A more elegant approach is to approximate any 3D geometry as
a composite created from a modest set of simpler generic shapes y = rz sin共␾兲 共5兲
共e.g., cylinder, cone, sphere, and plane兲. Each such primitive is where ␹ is a random number between 0 and 1, rz is the radius of
described not in terms of a meshing arrangement but rather by a the cylinder at z 共here rz = 1兲, and ␾ is a random angle between 0
set of surface equations. Any primitive may be manipulated via and 2p.
affine transformations that allow it to be scaled, rotated, and trans- However, for a tapered cylinder, it is no longer possible to
posed in 3D space to provide an infinite set of options while simply use a random number to represent the z coordinate but
requiring a computational method for just a single shape. Thus rather a random surface area fraction 共0 ⱕ A f ⱕ 1兲 is selected. As
any complex 3D geometry may be constructed as a collection of the base radius and height for this primitive are both unity, A f can
affine transformed primitives with such an approach being widely be expressed as the following function of z:
used in the computer-based multimedia industry as previously
noted 关3,4兴. z共1 + rz兲
Intuitively, it might be expected that rays are created and traced Af = 共6兲
共1 + s兲
in “world space,” that is, the space where an actual object is
approximated by a composite of 共affine transformed兲 3D primi- where rz = 1 + 共s − 1兲z is the radius of the tapered cylinder at the
tives. However, such an approach increases the complexity of the point z.
ray-tracing implementation as it necessitates calculation of each Substituting for rz in Eq. 共6兲 gives a quadratic equation whose
ray’s starting point and directional vector within the transformed positive solution gives a random z coordinate distribution which
space. It is computationally far more efficient to calculate a ray’s reflects the changing surface area of the tapered cylinder:
starting point and launch direction in the primitive space, and
subsequently transform the ray to perform the appropriate calcu-
冑1 − A f + A2f − 1
z= 共7兲
lations. This ray transformation procedure is shown schematically s−1
in Fig. 2 for the case where the objective is to determine whether The coordinates defining the starting point of any ray on this
a ray from one object 共A兲 intersects with a second object 共B兲. tapered cylinder are thus
In this way, it is possible to describe a wide variety of 3D
geometries in terms of a small number of primitive shapes and to rz = 1 + 共s − 1兲z 共8兲
use these compact descriptions to calculate the required view fac-
tors. For example, a polymer preform that is being drawn to fiber x = rz cos共␾兲 共9兲

072702-2 / Vol. 132, JULY 2010 Transactions of the ASME

Downloaded 29 Apr 2010 to 129.78.64.102. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
fiber in a relatively low temperature furnace where heat transfer to
the preform occurs by both convection and radiation. A critical
issue in mPOF fabrication is the extent of deformation undergone
by the hole structure as the preform is heated 共above the poly-
mer’s glass transition temperature兲 and drawn to fiber. Details on
mPOF fabrication are provided in the recent literature 关17兴.
In Fig. 3, the preform enters at the top of the furnace through an
adjustable iris while drawn fiber leaves at the bottom through a
second iris. Six external halogen lamps 共under on-off control兲 pro-
vide radiant heating through a central quartz “hot-zone” window
with the rest of the furnace being well insulated. The preform
surface is heated by thermal radiation 共both through the quartz
window and by reradiation from the furnace walls兲 and by con-
vection induced within the furnace. In the present view factor
determination context, two cases were considered. In the first, the
furnace was operated such that the preform never exceeded its
glass transition temperature and thus never underwent any neck-
Fig. 3 Schematic diagram of fiber drawing furnace down in diameter. Here all relevant view factors were determined
by numerical integration of Eq. 共1兲. As noted, very good agree-
ment was obtained between experimental temperature measure-
ments and results from a fully conjugate heat transfer model
y = rz sin共␾兲 共10兲
where convective heat transfer was estimated using commercial
The second issue is calculation of the ray directional vector, which computational fluid dynamics software.
is achieved by assuming azimuthal symmetry and calculating a In the second case considered 共for which experimental mea-
hemispherical distribution as suggested by Farmer and Howell surements were not available兲, it was assumed that the preform
关12兴. This distribution is rotated in such a way that it lies parallel was raised above its glass transition temperature and that the re-
to the surface of the tapered cylinder. The angle 共␪兲 by which the sultant preform neckdown could be described by the “slender
directional distribution needs to be rotated around the x-axis is body” approximation 关18兴. Here the preform radius 共r兲 as a func-
simply given by tion of distance from the furnace entrance 共z兲 is given by
␪ = tan−1共1 − s兲 共11兲
Applying this transformation and rotating around the z-axis by the
random angle ␾, a ray’s directional vector may be determined as
r = R exp 冉 z ln共Dr兲
2H
冊 共14兲
follows:
where Dr is the draw ratio 共set by furnace operating conditions兲, H

冤 冥冤 冥
xdir x cos共␾兲 − sin共␾兲 ⫻ 共y cos共␪兲 − z sin共␪兲兲 is the height of the furnace 共0.18 m兲, and R is the initial preform
y dir = x sin共␾兲 + cos共␾兲 ⫻ 共y cos共␪兲 − z sin共␪兲兲 共12兲 radius 共0.006 m兲.
zdir y sin共␪兲 − z cos共␪兲 In the latter case, determining all relevant view factors by nu-
merical integration would be a time-consuming exercise. How-
It should be noted that this method generates a uniform directional ever, calculating these view factors using ray tracing is scarcely
distribution and thus is unsuitable for cases involving materials more involved than for the case where the preform diameter re-
whose emissive characteristics deviate significantly from the un- mains unchanged. The two Dr values chosen 共9 and 900兲 gave exit
derlying gray body assumption. to inlet radius ratios 共i.e., r / R兲 that ranged from a modest to a
Using a ray’s starting point and directional vector, as defined by substantial preform neckdown 共33.33% and 3.33%, respectively兲.
Eqs. 共9兲, 共10兲, and 共12兲, ensures that the calculated view factors
are unbiased by any shape change, which in the present case 3.1 Constant Diameter Preform. Although not required by
would occur as the preform tapers and is drawn to fiber. this implementation of ray tracing, the preform and furnace walls
In applying the ray-tracing approach, a key computational pa- were each discretized using 180 “slices” to allow a direct com-
rameter is the ray density ␳ 共i.e., the rays launched per unit area兲 parison of the view factors with those obtained via numerical
to be used. Once this is set, the total number of rays used is simply integration 共which used this level of discretization to ensure that
the product of the surface area of the transformed 共into “world the results were grid independent兲. Additionally, ray densities 共␳兲
space”兲 tapered cylinder and the chosen ray density. Noting that of 104 and 105 per unit area were used to assess the impact of this
rotation and translation do not affect surface area, the following key parameter on the ray-tracing results. As expected, increasing
formula can be used to give the surface area of a tapered cylinder the ray density improved the “accuracy” of the ray-tracing results
that has been transformed by scaling factors a and b in the x and with excellent agreement between the two methods being
y directions, respectively: achieved at the higher ␳ value.


Results are given in Figs. 4–6 where in each case the view
␲共1 + s兲 s共a2 + b2兲 factors calculated by numerical integration and ray tracing are
A= 共13兲
s 2 shown as solid lines and discrete symbols, respectively. Although
all possible view factors for this furnace enclosure were deter-
mined, values are only given here between the preform surface
3 View Factor Results for Fiber-Drawing Furnace and the furnace wall 共observed from three different axial positions
In order to evaluate the performance of ray tracing, it was used z = 0.0095, 0.0895, and 0.1695 of one surface to the entire other
to determine the view factors needed in modeling the heat transfer surface or to itself兲 to provide a direct comparison between the
within a fiber drawing furnace. Figure 3 is a schematic represen- two methods. Here F p-f refers to the view factor from a position
tation of an operational furnace used to fabricate microstructured on the preform surface to the furnace wall, F f-f refers to the view
polymer optical fibers 共mPOFs兲. Such fibers owe their light- factor from a position on the furnace wall to itself, and so on. The
guiding properties to a pattern of holes that runs the length of the z values chosen show the essential geometric symmetry of a sys-
fiber. Here the required pattern is created by drilling holes into a tem that comprises a constant diameter cylinder located along
monolithic block referred to as the preform. This is then drawn to the axis of a cylindrical enclosure. Under conditions where the

Journal of Heat Transfer JULY 2010, Vol. 132 / 072702-3

Downloaded 29 Apr 2010 to 129.78.64.102. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Fig. 6 View factors from three positions „zf = 0.0095,
0.0895, 0.1695… on the furnace wall to the preform surface „left…
and itself „right… „␳ = 105 per unit area; cylindrical preform…

Fig. 4 View factors from three positions „zp = 0.0095,


0.0895, 0.1695… on the preform to the furnace wall „␳ = 104 per
unit area; cylindrical preform… constant emissivity ew of 0.885, a height H of 180 mm, and a
radius Rw of 32 mm; uniform wall temperature Tw of 373 K兲. Note
that here the net heat flux for the preform has been scaled by the
preform undergoes neckdown, however, this symmetry is broken, emitted flux qb from a black surface with the same uniform wall
as discussed in Sec. 3.2. temperature, that is, by qb = ␴0T4w, while the axial length has been
The accuracy of the view factor calculations is implied by both scaled by H. As expected with an aspect ratio for the cylindrical
the close agreement between all Fi-j profiles obtained from the preform of 30, when both irises are treated as insulated surfaces,
numerical integration and ray-tracing methods and the “sum rule” the net radiative heat flux over the middle portion of the preform
where the total of the view factors for each surface within the surface is consistent with the analytical result for two infinitely
enclosure is essentially unity. Using these view factors, it is pos- long concentric cylinders. However, if both irises are treated as
sible to determine the radiative heat transfer, using the net radia- black surfaces, then, due to heat lose thought the irises, the overall
tion method 关2,19兴, between a stationary, finite length, solid pre- net radiative heat flux to the preform surface is reduced along the
form, and the furnace surfaces where the heating wall is at a length of the preform, with the reduction becoming more marked
uniform temperature, and the top and bottom irises are either in- when approaching the ends of the preform.
sulated or treated as black surfaces. For this case, if the aspect 3.2 Reducing Diameter Preform. A constant diameter pre-
ratio of the cylindrical preform is large enough, the net radiative
form may be thought of as having a draw ration 共Dr兲 of unity. As
heat flux profile over the central section of the preform surface
Dr is increased, so the amount of preform neckdown increases.
should approximate that for the limiting case of two infinitely long
View factors within the drawing furnace were calculated for two
concentric cylinders. Figure 7 shows the results when a stationary
Dr values 共9 and 900兲 using a ray density of 105 per unit area. The
polymethylmethacrylate 共PMMA兲 preform 共R = 6 mm with a con-
stant hemispherical total emissivity e p of 0.90; initially at a uni-
form temperature of 293 K兲 is heated inside a furnace 共with a

Fig. 7 Net radiative heat flux profiles along a stationary cylin-


Fig. 5 View factors from three positions „zp = 0.0095, drical preform within a furnace with a uniform heating wall tem-
0.0895, 0.1695… on the preform to the furnace wall „␳ = 105 per perature; two different thermal boundary conditions are used
unit area; cylindrical preform… for the irises

072702-4 / Vol. 132, JULY 2010 Transactions of the ASME

Downloaded 29 Apr 2010 to 129.78.64.102. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Fig. 10 View factors from three positions „zf = 0.0095,
0.0895, 0.1695… on the furnace wall to the preform and to itself
Fig. 8 View factors from three positions „zp = 0.0095, „␳ = 105 per unit area; Dr = 9…
0.0895, 0.1695… on the preform to the furnace wall „␳ = 105 per
unit area; Dr = 9…

and a ray density 共105 per unit area兲 that provided acceptable
profiles presented in Figs. 8–11 correspond to those given in Figs. accuracy. This difference must, however, be weighted against the
5 and 6 for the constant diameter case. considerable time 共several man months兲 taken in carrying out the
Figures 5, 8, and 9 show the view factor profiles from the same necessary analysis and coding for the numerical integration case
positions on the preform surface to the furnace wall for three 关20兴, the fact that it was not felt to be cost effective to extend this
different Dr values 共1, 9, and 900兲. The amount of preform thin- approach to the general deforming preform case, and the very
ning increases with Dr and the form of the view factor profile modest time taken to set up a new case using the ray-tracing
changes as the system geometry becomes less longitudinally sym- approach. No attempt was made in this study to speed up the
metrical. Similar changes with preform thinning are apparent in ray-tracing calculations using either faster hardware or more so-
Figs. 6, 10, and 11, where view factor profiles from three different phisticated numerical techniques. In the latter category, con-
positions on the furnace wall to both the preform surface 共F f-p兲 strained maximum likelihood smoothing appears to have promise
and itself 共F f-f 兲 are given. 关21兴, although its implementation would result in a significant
nonlinear programming problem, somewhat at odds with the rela-
3.3 Computational Efficiency. It took ⬃5 s 共using an AMD tive simplicity of the coding required for ray tracing built around
Althon 64 processor, 2.21 GHz, 2.0 Gbytes of RAM兲 to perform a the use of primitives.
complete set of view factor calculations for the draw furnace with Finally, it should be noted that if numerical integration is used,
a constant diameter preform using numerical integration at a level then the extent of the geometric discretization is primarily set by
of discretization necessary to ensure grid independence. By com- the need to ensure that the computed view factors are grid inde-
parison, the view factor profiles for a slender body preform with a pendent. By comparison, the discretization used in ray tracing is
draw ratio of 900 took 44.2 h when run on a comparable computer set simply by the resolution required in any given view factor
profile. Thus in the present calculations, 180 cylindrical slices

Fig. 9 View factors from three positions „zp = 0.0095, Fig. 11 View factors from three positions „zf = 0.0095,
0.0895, 0.1695… on the preform to the furnace wall „␳ = 105 per 0.0895, 0.1695… on the furnace wall to the preform and to itself
unit area; Dr = 900…. „␳ = 105 per unit area; Dr = 900…

Journal of Heat Transfer JULY 2010, Vol. 132 / 072702-5

Downloaded 29 Apr 2010 to 129.78.64.102. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
共either constant or tapered兲 were used to describe the preform 关7兴 Steward, F. R., and Cannon, P., 1971, “The Calculation of Radiative Flux in a
Cylindrical Furnace Using the Monte Carlo Method,” Int. J. Heat Mass Trans-
simply for ease of comparison with the numerical integration re-
fer, 14, pp. 245–262.
sults. 关8兴 Fan, T., and Fedorov, A. G., 2002, “Apparent Radiative Properties and Radia-
tion Scattering by a Semi-Transparent Hemispherical Shell,” ASME J. Heat
4 Conclusions Transfer, 124, pp. 1088–1094.
关9兴 Hu, L., Schmidt, A., Narayanaswamy, A., and Chen, G., 2004, “Effect of
The determination of view factors for complex 3D geometries Periodic Structures on Coherence Properties of Blackbody Radiation,” ASME
can be a challenge. In some cases, especially those that exhibit a J. Heat Transfer, 126, pp. 786–792.
measure of geometric symmetry, it is possible to employ numeri- 关10兴 Xia, X. L., Shuai, Y., and Tan, H. P., 2005, “Calculation Techniques With the
cal integration techniques, although considerable analysis is gen- Monte Carlo Method in Stray Radiation Evaluation,” J. Quant. Spectrosc. Ra-
erally required and care needs to be taken in the discretization diat. Transf., 95, pp. 101–111.
关11兴 Vujičić, M. R., Lavery, N. P., and Brown, S. G. R., 2006, “View Factor
employed. Ray tracing is an alternative numerical approach that Method Using the Monte Carlo Method and View Factor Sensitivity,” Com-
does not rely on any detailed geometric analysis. In the version mun. Numer. Methods Eng., 22, pp. 197–203.
employed in this study, 3D geometries are constructed using a set 关12兴 Farmer, J. T., and Howell, J. R., 1998, “Comparison of Monte Carlo Strategies
of primitive shapes. A subsequent Monte Carlo simulation 共in for Radiative Transfer in Participating Media,” Adv. Heat Transfer, 31, pp.
which the “fate” of randomized rays leaving a surface are col- 333–429.
关13兴 Howell, J. R., 1968, “Application of Monte Carlo to Heat Transfer Problems,”
lated兲 is used to determine the required view factors. Although run
Adv. Heat Transfer, 5, pp. 1–52.
times for ray-tracing calculations are significantly longer than for 关14兴 Campbell, P. M., 1967, “Monte Carlo Method for Radiative Transfer,” Int. J.
numerical integration, the “setup” time for a problem using ray- Heat Mass Transfer, 10, pp. 519–527.
tracing is essentially independent of geometric complexity, mak- 关15兴 Vueghs, P., de Koning, H. P., Pin, O., and Beckers, P., 2009, “Use of Geometry
ing this a highly attractive approach for generic view factor soft- in Finite Element Thermal Radiation Combined With Ray Tracing,” J. Com-
put. Appl. Math. 共in press兲.
ware.
关16兴 Sakai, S., and Maruyama, S., 2003, “A Fast Approximated Method for Radia-
tive Exchange for Combined Heat Transfer Simulation,” Numer. Heat Trans-
References fer, Part B, 44, pp. 473–487.
关17兴 Large, M. C. J., Poladian, L., Barton, G. W., and van Eijkelenborg, M. A.,
关1兴 Howell, J. R., 2008, “A Catalog of Radiation Heat Transfer Configuration
2008, Microstructured Polymer Optical Fiber, Springer, New York.
Factors,” 2nd ed., http://www.me.utexas.edu/~howell/index.html
关18兴 Xue, S.-C., Tanner, R. I., Barton, G. W., Lwin, R., Large, M. C. J., and
关2兴 Siegel, R., and Howell, J. R., 2001, Thermal Radiation Heat Transfer, 4th ed.,
Poladian, L., 2005, “Fabrication of Microstructured Optical Fibers, Part I:
Taylor & Francis, New York.
关3兴 Hill, F. S., 2001, Computer Graphics Using Open GL, 2nd ed., P. Hall, ed., Problem Formulation and Numerical Modeling of Transient Draw Process,” J.
Prentice-Hall, Englewood Cliffs, NJ. Lightwave Technol., 23共7兲, pp. 2245–2254.
关4兴 Shirley, P., Sung, K., Brunvand, E., Davis, A., Parker, S., and Boulos, S., 2008, 关19兴 Xue, S.-C., Poladian, L., Barton, G. W., and Large, M. C. J., 2006, “Radiative
“Fast Ray-Tracing and the Potential on Graphics and Gaming Courses,” Com- Heat Transfer in Preforms for Microstructured Optical Fibres,” Int. J. Heat
put. Graph., 32, pp. 260–267. Mass Transfer, 50共7–8兲, pp. 1569–1579.
关5兴 Farmer, J., 1995, “Improved Algorithms for Monte Carlo Analysis of Radiative 关20兴 Lee, S. H.-K., and Jaluria, Y., 1995, “The Effects of Geometry and Tempera-
Heat Transfer in Complex Participating Medium,” Ph.D. thesis, University of ture Variations on the Radiative Transport During Optical Fiber Drawing,” J.
Texas, Austin, TX. Mater. Process. Manuf. Sci., 3, pp. 317–331.
关6兴 Wang, A., and Modest, M. F., 2006, “Photon Monte Carlo Simulation for 关21兴 Daun, K. J., Morton, D. P., and Howell, J. R., 2005, “Smoothing Monte Carlo
Radiative Transfer in Gaseous Media Represented by Discrete Particle Fields,” Exchange Factors Through Constrained Maximum Likelihood Estimation,”
ASME J. Heat Transfer, 128, pp. 1041–1049. ASME J. Heat Transfer, 127, pp. 1124–1128.

072702-6 / Vol. 132, JULY 2010 Transactions of the ASME

Downloaded 29 Apr 2010 to 129.78.64.102. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
View publication stats

You might also like