Sie sind auf Seite 1von 21

A Numerical Model for Radiative Heat Transfer Analysis in Arbitrarily-Shaped Axisymmetric Enclosures with Gaseous Media

(enunes@merlin.cc.manhattan.edu) and (naraghi@merlin.cc.manhattan.edu) Department of Mechanical Engineering Manhattan College Riverdale, NY 10471

Edmundo M. Nunes

Mohammad H.N. Naraghi

Abstract
A numerical model for evaluating thermal radiative transport in irregularly-shaped axisymmetric enclosures containing a homogeneous, isotropically scattering medium is presented. Based on the Discrete Exchange Factor (DEF) method, exchange factors between arbitrarily-oriented dierential surface/volume ring elements are calculated using a simple approach. The present method is capable of addressing blockage eects produced by inner/outer obstructing bodies. The results obtained via the current method are found to be in excellent agreement with existing solutions to several cylindrical media benchmark problems. The solutions to several rocket-nozzle and plug-chamber geometries are presented for a host of geometric conditions and optical thicknesses.

Nomenclature
a dA dr ds dsi sj dss DSi Sj DSS dsi vj dsv DSi Vj DSV dV dvi sj dvs DVi Sj DVS dvi vj dvv amplitude of cosine function in Eqs. (29) and (30) surface area of dierential ring element radial thickness of dierential volume ring element width of dierential surface ring element direct exchange factor between surface ring elements i and j surface-to-surface direct exchange factor matrix total exchange factor between surface ring elements i and j surface-to-surface total exchange factor matrix direct exchange factor between surface ring element i and volume ring element j surface-to-volume direct exchange factor matrix total exchange factor between surface ring element i and volume ring element j surface-to-volume total exchange factor matrix volume of dierential ring element direct exchange factor between volume ring element i and surface ring element j volume-to-surface direct exchange factor matrix total exchange factor between volume ring element i and surface ring element j volume-to-surface total exchange factor matrix direct exchange factor between volume ring elements i and j volume-to-volume direct exchange factor matrix 1

DVi Vj DVV dz E I Kt N q00 r ro r Ri (z) Ro (z) T w W z

total exchange factor between volume ring elements i and j volume-to-volume total exchange factor matrix axial thickness of dierential volume ring element emissive power identity matrix extinction coecient number of dierential ring elements radiative heat ux radial coordinate radius of base of enclosure position vector local radius of inner obstructing axisymmetric body local radius of outer obstructing axisymmetric body temperature numerical integration weight factor weight factor matrix axial coordinate

Greek Symbols i , j i , o absorption matrix angle between surface normal and vector connecting ri and rj emissivity function dened by Eq. (12) limiting values of cos due to blockage by inner/outer axisymmetric bodies tilt angle of surface with respect to z-axis azimuth angle reectivity reectivity matrix Stefan-Boltzmann constant = 5.67051 1008 W=(m2 K 4 ) transmittance ( = Kt ro ) 2

!o

scattering albedo

Subscripts i j max min s v w designates emitting ring element designates receiving ring element designates maximum designates minimum designates surface designates volume designates wall

Introduction
Thermal radiative transport continues to emerge as an important energy transfer mechanism in a wide variety of practical systems. For example, in high-pressure spacecraft engines, combustion products can reach very high temperatures, rendering radiation a signicant mode of heat transfer. Moreover, thermal radiation heat transfer analyses are of primary interest in other specialized computational arenas, such as gas turbine, plug-chamber, fusion reactor, and crystal growth thermal/uid modeling. Many of the aforementioned engineering systems are generally axisymmetric in shape, thereby reducing a computationally exhaustive 3-D analysis to a two-dimensional procedure. However, due to the inherent complexities associated with radiative transfer calculations, such as the long distance nature of radiation (solid angle integration), dependence on orientation between participating elements, and functional dependence of radiative properties, radiative analyses are very often simplied. Consequently, simplication or neglect of radiative phenomenon in thermophysical models may inaccurately predict temperature and heat ux proles. This is very much the issue in crystal growth modeling, where thermal gradients and interface shapes, which directly aect crystal quality and process stability, are sensitive to radiative interactions between participating surfaces and/or media. The aim of this paper is to present a model for systematically evaluating radiative heat transfer in arbitrarily-shaped axisymmetric enclosures with participating media. 3

A close examination of the radiative heat transfer literature reveals that the analysis of axisymmetric domains is limited to cylindrical enclosures, the simplest of axisymmetric geometries. Several numerical techniques have been implemented to solve for radiative transport within cylindrical geometries, including the Discrete-Ordinates (S-N) method, the Spherical Harmonics (P-N) method, the Monte-Carlo method, the YIX quadrature, and the zonal method. There are also several exact analytic solutions available for cylindrical media. Dua and Cheng [1] obtained analytic expressions for nite and innite cylinders with a non-scattering medium, whereas Crosbie and Dougherty [2] examined absorbing, emitting, and scattering media within axially nite, but radially innite cylinders. The Discrete Ordinates method is used by Jamaluddin and Smith [3]. Although the Discrete Ordinates method is accurate and less memory intense than other methods, it suers from ray eects [4]. Menguc and Viskanta [5] used the Spherical Harmonics method (P1 and P3) to analyze cylindrical enclosures. The P-N methods require a high order of approximation to achieve accurate results, especially in optically thin regions, and consume large amounts of computation time and memory. Stewart and Cannon [6] used the Monte Carlo method, which is exible, but time-consuming and suers from inaccuracies due to statistical error. Albeit highly accurate, the zonal method, used by Hottel and Sarom [7] to study radiative transfer in cylindrical furnaces, is computationally intensive, requiring the evaluation of multiple spatial integrations for computing exchange factors. All in all, there exists a need for a computationally ecient and exible scheme for evaluating radiative transport in axisymmetric enclosures. A numerical model has been developed based on the Discrete Exchange Factor (DEF) method [8]. The DEF method, based on a point-to-point approach for radiative analyses of enclosures, has proven to be computationally advantageous over the zonal method (since multiple integrations aren't necessary) and more accurate in one-dimensional systems [8]. The numerical results for twodimensional and three-dimensional systems are in excellent agreement with other methods [9, 10]. Since radiative exchange is computed between nodal points, the DEF method lends itself to grid compatibility with nite dierence/element schemes for solving combined-mode heat transfer/uid ow problems [11, 12]. Exchange factors between dierential ring element pairs are computed by generalizing Modest's [13] model for view factors between dierential ring elements on concentric bodies to give the appropriate DEF expressions. Blockage eects produced by inner and/or outer bodies are accounted for in the present formulation.

Mathematical Formulation
Consider the arbitrarily-shaped axisymmetric enclosure shown in Fig. 1a. The enclosure, which is comprised of an inner and outer surface, contains a radiatively participating medium. For simplicity, all surfaces are assumed to be opaque, diuse and gray, and the medium is homogeneous and isotropically scattering. The DEF method computes radiative heat exchange between spatial locations in enclosures by considering four avenues of direct radiative exchange: surface-to-surface, surface-to-volume, volume-to-surface, and volume-to-volume. Integrating the nodal DEF equations, given in [8], about the circumferential direction for dierential elements i and j gives the following direct exchange factor expressions between axisymmetric surface/volume ring elements: 2rj dsj Z max cos i cos j eKt jri rj j dj min jri rj j2

dsi sj =

(1)

2Kt rj drj dzj Z max cos i eKt jri rj j dsi vj = dj min jri rj j2 dvi sj = rj dsj Z max cos j eKt jri rj j dj 2 min jri rj j2

(2)

(3)

Kt rj drj dzj Z max eKt jri rj j dvi vj = dj 2 min jri rj j2

(4)

where symmetry with respect to the azimuth angle has been incorporated; ri denotes the position at which radiation is emitted; rj denotes the position at which radiation is received; is the angle between the surface normal and the vector connecting ri and rj ; subscripts i and j denote emitting and receiving elements, respectively; dsj is the width of the dierential surface ring element; and min and max are the minimum and maximum azimuth angles through which ring element j is seen from a point on ring element i , respectively. The allowable range of is dictated by the orientation and relative position of the ring element pair and blockage eects produced by inner and/or outer obstructing bodies. Details concerning the determination of the limiting azimuth angles follow from Modest's [13] formulation and are subsequently presented. Geometric consideration of any ring element pair depicted in Fig. 1a. reveals:

2 jri rj j2 = ri2 + rj 2ri rj cos + (zj zi )2

(5)

and for surface ring elements:

jri rj j cos i = (ri rj cos ) cos i (zj zi ) sin i jri rj j cos j = (ri cos rj ) cos j + (zj zi ) sin j

(6)

(7)

where k is the angle, resting in the r z plane, measured from the z-axis, in the direction of increasing radius, onto the backside of the element k . For all inward facing surface ring elements, =2 3=2, and for all outward facing ring elements, =2 =2. Combining Eqs. (1-7) gives the resultant exchange factor expressions:
2 2ri rj cos i cos j dsj Z max (i cos ) (j cos ) eKt jri rj j dj min jri rj j4 2 2Kt rj cos i drj dzj Z max (i cos ) eKt jri rj j dj min jri rj j3

dsi sj =

(8)

dsi vj =

(9)

ri rj cos j dsj Z max (j cos ) eKt jri rj j dvi sj = dj 2 min jri rj j3 Kt rj drj dzj Z max eKt jri rj j dvi vj = dj 2 min jri rj j2 where: i = ri zj zi + tan i ; rj rj j = rj zi zj + tan j ri ri

(10)

(11)

(12)

The limiting angles min and max remain to be determined. The limiting azimuth angles for surface-to-surface exchange are governed by the conguration and orientation of both surface ring elements and shadowing produced by inner and/or outer blocking bodies. Radiative exchange among surface ring elements inherently satises the following condition:

cos k 0 ; k = i; j 6

(13)

Combining the geometric relations obtained in Eqs. (6) and (7) with the condition given above results in an expression (functions i and j in Eq. (12)) for the cosine of potential limiting azimuth angles for a surface ring element pair. If computed values i and j are in the interval of [-1,1], the ring element pair is mutually fully or partial visible. For i and j outside of [-1,1], and depending on the orientation of the ring elements (i and j ), the ring elements may be fully visible/invisible to each other. Take, for instance, two horizontal surface ring elements facing each other (i = =2 and j = 3=2). Both elements are fully visible to each other, yet values of i and j are both 1, outside the interval [-1,1]. Now, consider two horizontal surface ring elements facing in the same direction, with i = j = =2. Both elements clearly can not exchange radiant energy, yet i and j are, as in the example noted above, 1. Values of cos min and cos max are systematically determined by referring to Table 1, where potential azimuthal limiting factors are categorized for specic ranges of cos for each surface element. If, upon referring to Table 1, cos min < cos max for a ring element pair, the exchange factor is taken as zero. It is important to note that the determination of values for min and max for exchange among surface/volume, volume/surface, and volume/volume ring elements is performed in a similar way. The dierence rests in recognizing that volume ring elements have no orientation and radiate over a solid angle of 4. Thus, for volume ring element j , the function j is meaningless (since j does not exist). Consequently, the selection of cos min and cos max from Table 1 becomes a simplied process. It is possible that, in many instances, the view between a ring element pair is partially obstructed by an inner and/or outer blocking body. Inner and outer blockage angles, cos1 i and cos1 o , are evaluated by projecting a line from a point on an emitting ring element (denoted by subscript i) around the periphery of the blocking body at an axial position zk , such that zk is between zi and zj . The intersection point between the receiving ring element (denoted by subscript j ) and shadow produced by the blocking body at zk result in a potential minimum/maximum azimuth angle. This procedure is repeated for several values of zk and can be mathematically stated as:
2 2 R2 (zk ) (zj zi )2 ri (zj zk )2 rj (zk zi )2 i i = max 2ri rj (zk zi ) (zj zk ) 2 2 R2 (zk ) (zj zi )2 ri (zj zk )2 rj (zk zi )2 o = min o 2ri rj (zk zi ) (zj zk )

"

# #

(14)
zk 2(zi ;zj )

"

(15)

zk 2(zi ;zj )

For an inner cylindrical obstructing body, Eq. (15) reduces to: 7

R2 i = i ri rj

"

2 Ri 1 2 ri

2 Ri 1 2 rj

!# 1
2

(16)

In an eort to further clarify the selection process of min and max , consider again the arbitrarilyshaped axisymmetric enclosure shown in Fig. 1a. The lines of view in the r z plane, due to ring conguration and blockage, from which ring elements dAj and dVj are seen from a point on dAi are shown. These lines are projected onto the r plane via vertical and horizontal cuts through the axisymmetric bodies, depicted in Fig. 1b. Labels a i denote circumferential positions on dAj and dVj , dividing them into viewed and unviewed portions. Note that the cosines of both i and j are both less than zero. The view from dAi to dAj is aected by i , j , i , and o , producing unviewed arcs ab; ac; hi ; and ae on dAj . It is evident that cos min = o and cos max = i . If there were no obstructing bodies, then cos min = j and cos max = 1. Radiative exchange between elements dAi and dVj is inuenced by i , i , and o , yielding o and i as cos min and cos max , respectively. A non-obstructed view between the ring elements would give values of i and 1 for cos min and cos max . Figure 1c displays projected lines of sight for exchange between dVi , dAj , and dVj . Arcs ab; ad ; and fg represent unseen strips of dAj produced, respectively, by potential limiting values of cos of o , j , and i . The cosine of the minimum and maximum limiting angles are j and i . For radiative exchange between dVi and dVj , cos min and cos max are obtained from arcs ac and eg and are designated as o and i . The exchange factors calculated from Eqs. (8-11) must satisfy the conservation of energy equations. The discretized form of these equations are given by:
Ns X Nv X Nv X

wsj dsi sj +

wvj dsi vj = 1

(17)

j=1 Ns X j=1

j=1

wsj dvi sj +

wvj dvi vj = 1

(18)

j=1

where wsk (=dsk ) and wvk (=drk dzk ) are numerical integration weight factors for the discretized conservation equations. These equations imply that radiant heat emitted from a surface or volume element will be absorbed by other surface and volume elements in the enclosure. Since the exchange factors are evaluated numerically via a 10-pt Gaussian integration routine, normalization procedures are required so as to obey the conservation laws. In order to account for direct radiative transfer 8

between surface/volume elements and diuse multiple reections at boundaries and isotropic scattering of radiation beams, total exchange factors are evaluated using an explicit matrix formulation of the DEF method. Radiative analysis of multi-dimensional anisotropic scattering media can be performed using the N-bounce or source function variations of the DEF method [14, 15]. However, the scope of this work is limited to computing in an isotropically scattering medium. Total exchange factors are computed using the equations outlined below:

DSS = [I fdss + dsv !o Wv [I dvv !o Wv ]1 dvsg Ws ]1 fdss + dsv !o Wv [I dvv !o Wv ]1 dvsg (19)

DSV = [I fdss + dsv !o Wv [I dvv !o Wv ]1 dvsg Ws ]1 dsv[I !o Wv dvv]1 (1 !o ) (20)

DVS = [I dvv !o Wv ]1 dvs [I Ws fdss + dsv !o Wv [I dvv !o Wv ]1 dvsg]1 (21)

DVV = [I dvv !o Wv ]1 dvv (1 !o ) + [I dvv !o Wv ]1 dvs Ws [I fdss + dsv !o Wv [I dvv !o Wv ]1 dvsg Ws ]1 dsv [I !o Wv dvv]1 (1 !o ) (22)

where DSS = [DSi Sj ], DSV = [DSi Vj ], DVS = [DVi Sj ], DVV = [DVi Vj ] are matrices of total exchange factors between dierential surface/volume axisymmetric ring elements; dss = [dsi sj ], dsv = [dsi vj ], dvs = [dvi sj ], dvv = [dvi vj ] are matrices of direct exchange factors between dierential surface/volume ring elements; Ws = [ws;i i;j ] and Wv = [wv;i i;j ] are diagonal matrices of numerical integration weight factors for surface/volume ring elements, respectively; = [ i;j ] and = [ i;j ] are diagonal matrices of reectivities and absorptivities for surface ring elements. Like

the direct exchange factors, the total exchange factors must obey conservation laws similar to those given in Eqs. (17) and (18).
Ns X Ns X Nv X

wsj DSi Sj +

wvj DSi Vj = 1

(23)

j=1

j=1

wsj DVi Sj +

j=1

j=1

Nv X

wvj DVi Vj = 1

(24)

The radiative heat ux at each surface/volume ring element is computed using the following energy balance equation:
00 qs;i = Es;i Ns X Ns X Nv X Nv X

j=1

ws;j DSj Si Es;j ws;j DSj Vi Es;j

wv;j DVj Si Ev;j

(25)

j=1

000 qv;i = Ev;i

wv;j DVj Vi Ev;j

(26)

j=1

j=1

where the gray emissive powers of the surface/volume elements, Es;i and Ev;i are dened as:
4 Es;i = i Ts;i

(27)

4 Ev;i = 4Kt (1 !o ) Tv;i

(28)

Results and Discussion


All numerical simulations were performed on a Silicon Graphics workstation with a computational time of approximately 20.3 seconds for a typical 20 10 mesh. All simulations were additionally performed for a 10 5 and 40 20 mesh, each consuming 0.94 and 945.3 seconds of computer time, respectively, to ensure grid independence of the presented solutions. Several benchmark problems were identied and solved in an eort to validate the model presented here. The rst benchmark problem concerns a cold, black cylinder with equivalent height and diameter, encompassing a non-scattering, uniform temperature medium. Although Kassemi and Naraghi [16] have solved this problem using the DEF method, we use it to verify the current formulation/code in the absence of other benchmark data. Figure 2 presents the dimenionsless radiative wall ux prole 10

for several solution techniques including an exact solution [1], spherical harmonics (P3) solution [5], YIX solution [17], and the present DEF solution. The DEF solution, similar to the YIX results, is in excellent agreement with the exact solution for all levels of optical thickness ( = 0.1, 1.0, and 5.0) shown. The P3 solution, however, deviates from the exact solution for low optical thicknesses ( 1.0), overpredicting the heat ux at the cylinder ends. There are other cylindrical media benchmark data available for comparison. Wu and Fricker [18] have compiled experimental wall heat ux data for a Delft furnace. The furnace, measuring 0.90 m in diameter and 5 m in length, contains a gray, non-scattering medium with an average extinction coecient of 0.3 m1. Table 2 gives the non-uniform temperature distribution within the gas. The furnace walls are gray and diuse with an emissivity of 0.8 and constant temperature of 425 K. The computed heat ux distributions for the Discrete-Ordinates [3], YIX/16 [17], Finite Volume [19], P3 [5] and DEF methods are presented in Fig. 3. All of the solution methods, evidently predict the location of the peak heat ux accurately. In fact, most of the solutions are moderately close to one another, with the exception of the P3 method, which seriously under-predicts the maximum heat ux value. There are, however, some note-worthy dierences between the experimental data and numerical solutions. These dierences are conceivably due to the presumption of a homogeneous medium [17]. In high-pressure rocket-nozzles, comprehensive radiative analyses are dicult to perform due to the complexities introduced by shadowing eects at the throat. Hammad and Naraghi [20] have developed a one-dimensional DEF scheme for evaluating radiant heat uxes in rocket-engine geometries. In their formulation, the exchange factors are computed between dierential volume plates and/or surface ring elements, warranted by the negligible variation of the combustion product properties in the radial and azimuthal directions. Results obtained by the 1-D [20] and 2-D (present method) are compared in Fig. 4a for a typical rocket-engine with radial/axial coordinates and surface/gas element temperatures specied in Table 3. The nozzle wall is gray and diuse with an emissivity of 0.95; the combustion gases are gray and non-scattering. The entrance/exit sections of the nozzle, shown at the left-hand/right-hand sides of the computational mesh in Fig. 4b, are assumed to be black, with temperatures xed at the rst surface segment and exit gas temperatures, respectively. The results show, that for extinction coecients of (Kt = 0.025 and 2.5 in1), both methods are in good agreement with one another. However, there are some discrepancies between the two methods for moderate extinction coecients (Kt = 0.25 in1 ). It is likely that 11

these dierences can be attributed to numerical inaccuracies associated with computing the exchange factors for the 1-D method. These exchange factors require the evaluation of double, triple, and quadruple integrals for computing surface-to-surface, surface-to-volume/volume-to-surface, and volume-to-volume exchange factors. For high optical thickness, the greatest contributor to the surface ux comes from gas elements that are very close to the surface. The computational error in such a case is limited to a few volume-to-surface/surface-to-volume exchange factor calculations. As the medium becomes less optically dense, all four modes of radiative exchange become signicant between all of the elements in the domain, resulting in a relatively large build up of numerical inaccuracies. As the medium becomes more and more optically thin, gas emission decreases signicantly, and radiative transport among surfaces dominates, requiring the evaluation of several double integrals. Given the conditions outlined in the rst benchmark problem, we have compiled the wall ux distribution for nozzle shapes generated using the function below: i r(z) = 1 a f(z) = 1 a 1 cos 2 Nw + 1

(29)

where a denotes the amplitude of the cosine function; Nw is the total number of ring elements comprising the wall, and i is the wall element number, between 1 and Nw . Several simulations were performed for nozzle shapes with amplitudes ranging in value from 0.05 to 0.45 in steps of 0.05 (see Fig. 5a for a sample mesh layout). For a 0.25, the wall ux proles, given in Figs. 6 - 8 for media with = 0.1, 1.0 and 5.0, show that the heat ux continually decreases from a peak value at the throat (z=ro = 1) to a minimum at the cylinder ends (z=ro = 2). As the amplitude of the wall generating function is increased beyond 0.25, the throat becomes more narrow. This, in turn, makes it dicult for surface elements in the vicinity of the throat (approximately 1.0 z=ro 1.4) to view the entire domain. Consequently, the heat ux is relatively low at the throat and depending on the value of a, peaks at approximately 1.3 z=ro 1.4 for media with low-to-moderate optical thicknesses ( = 0.1 and 1.0). Similar trends are noted for optically thick media ( = 5.0), due to the relative low emission of radiant energy from volume elements in the throat region. Radiative analyses of the aforementioned axisymmetric systems have required, at most, consideration of blockage eects produced by outer obstructing bodies. There are, however, many cases of practical importance where shadowing eects are caused by inner axisymmetric bodies. In 12

Czochralski growth systems, for example, a crystal pulled from the melt partially obstructs the view between the furnace walls and melt surfaces. In the area of spacecraft propulsion, the signicant cost advantages of testing rocket nozzles by using plug-chambers is sparking the attention of many researchers. The plug-chamber, used to simulate the ow-physics of rocket nozzles, consists of a diverging-converging inner plug contained within an outer cylindrical wall (see Fig. 5b). Several analyses were performed on various plug-chamber shapes generated using the following function: i r(z) = b + a f(z) = b + a 1 cos 2 Nw + 1

(30)

where b is the radius of the plug at the cylinder base and a is the amplitude of the cosine function. Figures 9 - 11 give the wall ux proles for numerous plug-chambers shapes (see Fig. 5c for a sample mesh layout) with a = 0.0 (inner obstructing cylinder case) to 0.40 in increments of 0.05, b = 0.1, and = 0.1, 1.0, and 5.0. The thermal conditions are equivalent to those prescribed in the rst benchmark problem. The gures exhibit trends similar to those reported for the rocket-nozzle shapes. For a 0.20, the dimensionless wall ux decreases smoothly from a maximum at z=ro = 1 to a minimum at z=ro = 2. As the value of a is increased further, the heat ux at the middle of the cylinder (z=ro = 1) drops due to the dimishing visibility of this region by the computational domain, shifting the peak ux to a position given by 1.6 z=ro 1.8. It should be noted that for = 5.0, the wall ux distributions for all of the plug-chamber shapes examined converge in the region near the cylinder ends. This phenemonon is explainable in light of the short-distance a radiative beam may travel in an optically dense medium and the xed conguration of the cylindrical wall.

Concluding Remarks
The numerical evaluation of radiative heat transfer in arbitrarily-shaped axisymmetric enclosures can be performed easily with the model presented in this work. The DEF method, together with a generalized view factor formulation between arbitrarily-oriented dierential ring elements form the basis for systematically evaluating direct exchange factors between surface and/or volume elements. Since nodal placement is quite arbitrary and exchange factors are evaluated in a direct manner, the code is ideal for modeling radiative phenomena in axiymmetric, multi-phase thermaluid systems with moving and/or deforming free-boundaries/interfaces. It is well-worth mentioning that the exchange factors between arbitrarily-oriented ring elements are evaluated via one numer-

13

ical integration, making the current method exible and computational ecient. A comparison of the results obtained for several cylindrical media benchmark problems illustrate the accuracy of the method for all levels of optical thickness. A comparison of 1-D and 2-D DEF solutions for a rocket engine demonstrate good agreement between both methods for low/high optical thicknesses. However, at intermediate optical conditions, the 2-D method is found to be more accurate, since multiple integrations (two-to-four per exchange factor) between all elements comprising the domain are not required. The solution to several nozzle and plug-chamber shape geometries are included to contribute to the benchmark literature.

Acknowledgments
This material is based upon work supported under a National Science Foundation Graduate Research Fellowship and by ARPA/AFOSR, as a part of The Consortium for Integrated Intelligent Modeling, Design, and Control of Crystal Growth Processes.

References
[1] Dua, S.S. and Cheng, P., Multi-Dimensional Radiative Transfer in Non-Isothermal Cylindrical Media With Non-Isothermal Bounding Walls, International Journal of Heat and Mass Transfer, Vol. 18, pp. 245-259, 1975. [2] Crosbie, A.L. and Dougherty, R.L., Two-Dimensional Radiative Transfer in a Cylindrical Geometry With Anisotropic Scattering, Journal of Quantitative Spectroscopy and Radiative Transfer, Vol. 25, pp. 551-562, 1981. [3] Jamaluddin, A.S. and Smith, P.J., Predicting Radiative Transfer in Axisymmetric Cylindrical Enclosures Using the Discrete Ordinates Method, Combustion Science Technology, Vol. 62, pp. 173-186, 1988. [4] Lathrop, K.D., Ray Eects in Discrete-Ordinates Equations, Nuclear Science Engineering, Vol. 32, pp. 357-369, 1968. [5] Menguc, M.P. and Viskanta, R., Radiative Heat Transfer in Axisymmetric Finite Cylindrical Enclosures, Journal of Heat Transfer, Trans. ASME, Vol. 108, pp. 271-276, 1986.

14

[6] Stewart, F.R. and Cannon, P., The Calculation of Radiative Heat Flux in a Cylindrical Furnace Using the Monte-Carlo Method, International Journal of Heat and Mass Transfer, Vol. 14, pp. 245-261, 1971. [7] Hottel, H.C. and Sarom, A.F., The Eect of Gas Flow Patterns on Radiative Transfer in Cylindrical Furnaces, International Journal of Heat and Mass Transfer, Vol. 8, pp. 1153-1169, 1965. [8] Naraghi, M.H.N., Chung, B.T.F., and Litkouhi, B., A Continuous Exchange Factor Method for Radiative Exchange in Enclosures With Participating Media, Journal of Heat Transfer, Vol. 110, No. 2, pp. 456-462, 1988. [9] Naraghi, M.H.N. and Kassemi, M., Radiative Transfer in Rectangular Enclosures: A Discretized Exchange Factor Solution, Journal of Heat Transfer, Trans. ASME, Vol. 111, No. 4, pp. 1117-1119, 1989. [10] Naraghi, M.H.N. and Litkouhi, B., Discrete-Exchange Factor Solution of Radiative Heat Transfer in Three-Dimensional Enclosures, Radiation Heat Transfer: Fundamentals and Applications, ASME publication HTD-Vol. 137, pp.133-140, 1989. [11] Saltiel, C. and Naraghi, M.H.N., Analysis of Radiative Heat Transfer in Participating Media Using Arbitrary Nodal Distribution, Numerical Heat Transfer Journal, Part B: Fundamentals, pp. 227-243, 1990. [12] Saltiel, C. and Naraghi, M.H.N., Combined-Mode Heat Transfer in Radiatively Participating Media Using the Discrete Exchange Factor Method with Finite Elements, Heat Transfer 1990, Hemisphere Publishing Corporation, Vol. 6, pp. 391-396, 1990. [13] Modest, M.F., Radiative Shape Factors Between Dierential Ring Elements on Concentric Axisymmetric Bodies, AIAA Journal of Thermophysics and Heat Transfer, Vol. 2, No. 1, pp. 86-88, 1988. [14] Naraghi, M.H.N. and Huan, J., An N-Bounce Method for Analysis of Radiative Transfer in Enclosures with Anisotropically Scattering Media, Journal of Heat Transfer, Vol. 113, No. 3, pp. 774-777, 1991.

15

[15] Huan, J. and Naraghi, M.H.N., Source Function Approach for Radiative Heat Transfer Analysis, AIAA Journal of Thermophysics and Heat Transfer, Vol. 6, No. 3, pp. 568-571, 1991. [16] Kassemi, K. and Naraghi, M.H.N., Application of Discrete Exchange Factor Method to Combined Heat Transfer Problems in Cylindrical Media, HT-6D: Transport Phenomena in Manufacturing and Materials Processing{II, Proc. 1996 ASME Winter Annual Meeting, Atlanta, Georgia, 1996. [17] Hsu, P.-F. and Ku, J.C., Radiative Heat Transfer in Finite Cylindrical Enclosures with NonHomogeneous Participating Media, AIAA Journal of Thermophysics and Heat Transfer, Vol. 8, No. 3, pp. 434-440, 1994. [18] Wu, H.L. and Fricker, M., The Characteristics of Swirl-Stabilized Natural Gas Flames{Part 2: The Behavior of Swirling Jet Flames in a Narrow Cylindrical Furnace, Journal Inst. of Fuel, Vol. 49, pp. 144-151, 1976. [19] Chui, E.H., Raithby, G.D., and Hughes, P.M.J., Prediction of Radiative Transfer in Cylindrical Enclosures with the Finite Volume Method, AIAA Journal of Thermophysics and Heat Transfer, Vol. 6, No. 4, pp. 605-611, 1992. [20] Hammad, K.J. and Naraghi, M.H.N., Exchange Factor Model for Radiative Heat Transfer Analysis in Rocket Engines, AIAA Journal of Thermophysics and Heat Transfer, Vol. 5, No. 3, pp. 327-334, 1991.

16

Table 1: Limiting values for the cosine of the azimuth angle. cos i 0 cos j 0 cos min = min(o ; 1) cos max = max(i ; j ; i ; 1) cos j 0 cos min = min(j ; o ; 1) cos max = max(i ; i ; 1) cos i 0 cos min = min(i ; o ; 1) cos max = max(j ; i ; 1) cos min = min(i ; j ; o ; 1) cos max = max(i ; 1)

17

Table 2: Temperature distribution within the Delft furnace [18]. z, (m) 0.15 0.45 0.75 1.05 1.35 1.65 1.95 2.25 2.55 2.85 3.15 3.45 3.75 4.05 4.35 4.65 4.95 T (r = 0:075m), (K) T (r = 0:225m), (K) T (r = 0:375m), (K) 1470 1120 870 1600 1320 1070 1620 1470 1360 1610 1550 1370 1580 1520 1350 1520 1470 1320 1470 1410 1280 1410 1360 1250 1350 1310 1210 1310 1260 1170 1270 1230 1150 1240 1200 1110 1200 1160 1090 1170 1130 1080 1140 1100 1070 1110 1080 1060 1080 1070 1060

18

Table 3: Nozzle radii and surface/gas temperature distribution along axial direction [20]. z, (in) -7.572 -6.843 -6.115 -5.386 -4.658 -3.929 -3.200 -2.472 -1.743 -1.015 -0.286 0.443 1.171 1.900 2.628 3.357 4.086 4.814 5.543 6.271 7.000 r, (in) Ts , (R) Tv , (R) 2.4199 1250.0 6708.6 2.3824 1244.4 6708.3 2.3374 1237.1 6707.5 2.2802 1222.0 6706.0 2.1951 1208.8 6703.6 2.0821 1205.1 6699.9 1.9446 1131.2 6692.6 1.7924 1062.9 6680.2 1.6380 1078.1 6660.8 1.4833 1106.5 6621.9 1.3743 1105.8 6509.7 1.4056 908.8 6150.6 1.5670 896.8 5885.3 1.7629 926.0 5692.7 1.9576 871.3 5549.5 2.1534 803.3 5415.1 2.3481 736.2 5306.3 2.5437 669.3 5203.9 2.7408 579.1 5116.1 2.9422 479.6 5034.3 3.0993 380.0 4952.5

19

Figure Captions
Figure 1: (a) Two-dimensional view of axisymmetric bodies in r z plane; (b) Projection of view from surface ring element dAi to ring elements dAj and dVj in r plane, and (c) Projection of view from volume ring element dVi to ring elements dAj and dVj in r plane. Figure 2: A comparison of dimensionless wall ux proles obtained from DEF, Exact, YIX, and P3 methods for cylindrical enclosures with = 0.1, 1.0 and 5.0. Figure 3: A comparison of wall heat ux distributions predicted using DEF, YIX, Finite Volume, S4 and P3 methods with experimental Delft furnace data [18]. Figure 4: (a) A comparison of dimensionless wall uxes based on 1-D [20] and 2-D (present) DEF methods for Kt = 0.025, 0.25 and 2.5 in1 and (b) Computational mesh of rocket engine. Figure 5: (a) Mesh layout for a nozzle with a = 0.25; (b) Geometric conguration of a sample plug-chamber, and (c) Mesh layout for a plug-chamber with a = 0.20. Figure 6: Dimensionless wall ux proles of nozzles with a ranging in value from 0.05 to 0.45 in steps of 0.05 for = 0.1. Figure 7: Dimensionless wall ux proles of nozzles with a ranging in value from 0.05 to 0.45 in steps of 0.05 for = 1.0. Figure 8: Dimensionless wall ux proles of nozzles with a ranging in value from 0.05 to 0.45 in steps of 0.05 for = 5.0. Figure 9: Dimensionless wall ux proles of plug-chambers with a ranging in value from 0.0 to 0.40 in steps of 0.05 for = 0.1. Figure 10: Dimensionless wall ux proles of plug-chambers with a ranging in value from 0.0 to 0.40 in steps of 0.05 for = 1.0. Figure 11: Dimensionless wall ux proles of plug-chambers with a ranging in value from 0.0 to 0.40 in steps of 0.05 for = 5.0. 20

Das könnte Ihnen auch gefallen