Você está na página 1de 90

COUPLED FLUID FLOW AND RADIATION MODELING OF A SMALL

PARTICLE SOLAR RECEIVER

_______________

A Thesis

Presented to the

Faculty of

San Diego State University

_______________

In Partial Fulfillment

of the Requirements for the Degree

Master of Science

in

Mechanical Engineering

_______________

by

Adam Winfield Crocker

Fall 2012
iii

Copyright 2012
by
Adam Winfield Crocker
All Rights Reserved
iv

ABSTRACT OF THE THESIS

Coupled Fluid Flow and Radiation Modeling of a Small Particle


Solar Receiver
by
Adam Winfield Crocker
Master of Science in Mechanical Engineering
San Diego State University, 2012

In recent years, concentrating solar thermal power has emerged as the most promising
technology for utility scale solar electricity generation. Central receiver systems, which are
one method of concentrated solar power, use a field of sun-tracking mirrors called heliostats
to focus light on a receiver. Existing receivers used in these systems have temperature and
flux limitations, which prevent the use of advanced power cycles and reduces plant
efficiencies compared to fossil fuel power plants. The development of air-cooled receivers
and small particle receivers in particular are summarized herein.
A new type of receiver has been proposed, which makes use of small carbon particles
and volumetric absorption in a gas-particle mixture to heat air directly. This thesis builds on
previous modeling work done in FORTRAN on the San Diego State University small particle
receiver project, expanding the Monte Carlo ray-trace model to include the computation fluid
dynamics capabilities of ANSYS FLUENT with the use of several user-defined functions.
The input flux is modified to more closely match that provided by a real heliostat field, and
the geometry is changed to more accurately approximate a real receiver.
The updated model is benchmarked against existing analytical solutions where
possible, and compared to the results of the previous model. The new model is run for a
variety of gas mass-flow rates, inlet power levels, and power distributions with a baseline
target input power of 5 MW. Outlet gas temperatures predicted by the model ranged from
1300 K to 1550 K, and receiver thermal efficiencies ranged from 80% to 91% depending on
operating conditions. The highest efficiencies predicted are with the highest mass-flow rate
tested of 6 kg/s.
v

TABLE OF CONTENTS

PAGE

ABSTRACT ............................................................................................................................. iv
LIST OF TABLES .................................................................................................................. vii
LIST OF FIGURES ............................................................................................................... viii
NOMENCLATURE ................................................................................................................ xi
ACKNOWLEDGEMENTS .................................................................................................... xii
CHAPTER
1 INTRODUCTION .........................................................................................................1
1.1 SPHER Receiver ................................................................................................9
1.2 Monte Carlo Ray-Trace Method ........................................................................9
1.3 Scope of Thesis ................................................................................................10
2 BACKGROUND ON SMALL PARTICLE RECEIVERS .........................................12
3 PREVIOUS WORK AND MODIFICATIONS TO MONTE CARLO RAY-
TRACE MODEL .........................................................................................................16
3.1 Incident Flux ....................................................................................................17
3.2 Outlet Tube ......................................................................................................20
3.3 Comparison to Benchmarks .............................................................................21
4 COUPLING OF THE MONTE CARLO AND FLUENT MODELS .........................29
4.1 Coupling Scheme & UDF Functions ...............................................................30
4.2 Solution Controls .............................................................................................34
4.3 Grid Matching Scheme ....................................................................................36
5 COLD FLOW & GRID STUDY .................................................................................40
5.1 Grid and Geometry ..........................................................................................40
5.2 Cold Flow Verification ....................................................................................43
5.3 Grid Study ........................................................................................................45
6 RESULTS AND DISCUSSION ..................................................................................54
6.1 Gaussian Flux Distribution ..............................................................................55
6.2 Flow Rate Variation .........................................................................................60
6.3 Input Power variation .......................................................................................65
vi

7 CONCLUSIONS AND FUTURE WORK ..................................................................70


7.1 Conclusions ......................................................................................................70
7.2 Future Work .....................................................................................................71
REFERENCES ........................................................................................................................74
APPENDIX
SUMMARY OF TURBULENCE MODEL CONSTANTS USED IN FLUENT.............77
vii

LIST OF TABLES

PAGE

Table 5.1. Summary of Grid Attributes ...................................................................................46


Table 6.1. Summary of Model Runs ........................................................................................54
viii

LIST OF FIGURES

PAGE

Figure 1.1. Schematic of a central receiver solar thermal power plant......................................1


Figure 1.2. Aerial view of the Gemasolar power plant in Spain................................................2
Figure 1.3. Rankine cycle diagram and schematic. ...................................................................3
Figure 1.4. Brayton cycle diagram and schematic. ....................................................................4
Figure 1.5. Combined Brayton-Rankine cycle diagram and schematic.....................................5
Figure 1.6. Diagram of the Solar One receiver. .........................................................................6
Figure 1.7. Schematic of pressurized high-temperature DLR receiver. ....................................7
Figure 1.8. Schematic of the Sandia falling particle receiver. ...................................................7
Figure 1.9. Schematic of cylindrical small particle solar receiver .............................................8
Figure 2.1. Cross section of the experimental SPHER ............................................................12
Figure 2.2. Schematic Cross section of the Weizmann Institute receiver. ..............................13
Figure 3.1. Temperature field output from original MCRT code. Flow is from right to
left, as indicated by white arrows. ...............................................................................18
Figure 3.2. Flux map predicted from MIRVAL. Figure represents 3 m by 3 m region
on aperture plane. .........................................................................................................19
Figure 3.3. Plot of the local incident solar flux applied to the aperture. ..................................20
Figure 3.4. Schematic of modified receiver geometry in Monte Carlo ray trace model .........21
Figure 3.5. Schematic of modified receiver geometry in Monte Carlo ray trace model
for black end-walls benchmark. ...................................................................................22
Figure 3.6. Net energy exchange between parallel black plates with no participating
medium, both at 500 K. The thick black line at 2.0 m represents the outlet
tube...............................................................................................................................22
Figure 3.7. Geometry of Monte Carlo model for concentric cylinders benchmark. ................23
Figure 3.8. Net energy exchange between concentric cylinders with no participating
medium. The inner cylinder is at 500 K and the outer cylinder is at 1000 K. .............24
Figure 3.9. Plot of non-dimensional emissive power in gray gas plane layer
benchmark. ...................................................................................................................25
Figure 3.10. Plot of non-dimensional emissive power in gray gas between concentric
cylinders. Each data set represents the radial temperature profile at a different
axial location. ...............................................................................................................27
ix

Figure 3.11. Plot of non-dimensional emissive power in gray gas between concentric
cylinders for three different optical depths. .................................................................28
Figure 4.1. Simple flow diagram of the coupling scheme used by the model. ........................30
Figure 4.2. Wall Flux Interpolation Method. Vertical grid lines correspond to wall
element divisions in the Monte Carlo solver. A negative net wall flux means
that the wall cell is a net emitter of radiation and thus is heated by the fluid,
while a positive value means the cell is a net absorber of radiation and is
cooled by the gas. .........................................................................................................32
Figure 4.3. Thermal boundary layers near the outer cylinder wall at four different
axial locations. Legend values indicate axial distance from aperture of each
data series. ....................................................................................................................35
Figure 4.4. Net energy gain in the gas flow as a function of model iteration number
for two different temperature under-relaxation methods. ............................................37
Figure 4.5. Net energy gain in the gas flow as a function of model iteration number
with no temperature under-relaxation. This scheme appears to be unstable, or
at least not converging. ................................................................................................38
Figure 4.6. Illustration of a FLUENT grid cell (in red) overhanging the boundary of a
Monte Carlo grid cell (in black)...................................................................................38
Figure 5.1. Schematic of the geometry as it is modeled in FLUENT and the Monte
Carlo code. Drawing is not to scale. ............................................................................40
Figure 5.2. View of the grid as it appears in FLUENT............................................................41
Figure 5.3. The FLUENT grid (green) with an approximate overlay of the Monte
Carlo grid (black). ........................................................................................................42
Figure 5.4. Plot of the velocity profile on the outlet plane of the receiver for two
FLUENT turbulence models as compared to a theoretical turbulent velocity
profile. ..........................................................................................................................44
Figure 5.5. Cold flow pressure contours in FLUENT .............................................................45
Figure 5.6. Wall temperatures in FLUENT for each of the three grids studied. .....................46
Figure 5.7. Velocity vectors in the window corner region for Grid 1. Vectors are
colored and scaled by magnitude. Color-bar scale represents velocity in m/s,
and length scale of the geometry is indicated at the top of the figure. ........................47
Figure 5.8. Velocity vectors in the window corner region for Grid 2. Vectors are
colored and scaled by magnitude. Color-bar scale represents velocity in m/s,
and length scale of the geometry is indicated at the bottom of the figure. ..................48
Figure 5.9. Velocity vectors in the window corner region for Grid 3. Vectors are
colored and scaled by magnitude. Color-bar scale represents velocity in m/s,
and length scale of the geometry is indicated at the top of the figure. ........................48
Figure 5.10. Cylinder wall temperatures in the first half-meter of the receiver. The x-
axis is scaled to align with the velocity vectors in Figure 5.11. ..................................49
x

Figure 5.11. Velocity vectors in the window corner region for Grid 2. Vectors are
scaled by magnitude of velocity, and colored by temperature in Kelvin.....................50
Figure 5.12. Wall fluxes imposed by the Monte Carlo on the FLUENT solver for each
of the three grids studied. Positive values of this net wall flux means the wall
is absorbing more radiation than it is emitting.............................................................52
Figure 6.1. Effect of solar flux distribution on energy balance. ..............................................56
Figure 6.2. Temperature field inside receiver with 5 MW of uniform incident solar
flux and 5 kg/s mass-flow rate. Color-map values are in units of Kelvin. ..................57
Figure 6.3. Temperature field inside receiver with 5 MW of Gaussian distributed
incident solar flux and 5 kg/s mass-flow rate. Color-map values are in units of
Kelvin...........................................................................................................................57
Figure 6.4. Cylinder wall temperature profiles for the uniform (blue) and Gaussian
(red) flux distribution cases. ........................................................................................59
Figure 6.5. Effect of mass-flow rate on gas outlet temperature and receiver thermal
efficiency, with 5 MW collimated Gaussian input. .....................................................60
Figure 6.6. Receiver, engine, and total efficiencies as a function of operating
temperature. .................................................................................................................62
Figure 6.7. Temperature field in the receiver with a mass-flow rate of 4.0 kg/s and
input power of 5 MW. Color-map numbers are in units of Kelvin..............................62
Figure 6.8. Temperature field in the receiver with a mass-flow rate of 6.0 kg/s and
input power of 5 MW. Color-map numbers are in units of Kelvin..............................63
Figure 6.9. Energy balance as a function of iteration number, for five different mass-
flow rate cases, 5 MW input power. ............................................................................64
Figure 6.10. Axial cylinder wall temperature profiles for five different mass-flow
rates, 5 MW input power. ............................................................................................65
Figure 6.11. Gas outlet temperatures and receiver thermal efficiencies for five
different inlet power levels. .........................................................................................66
Figure 6.12. Temperature field inside the receiver with a mass-flow rate of 2 kg/s and
input power of 2 MW. Color-bar units are Kelvin.......................................................67
Figure 6.13. Temperature field inside the receiver with a mass-flow rate of 6 kg/s and
input power of 6 MW. Color-bar units are Kelvin.......................................................68
Figure 6.14. Axial cylinder wall temperature profiles for five different inlet power
levels and mass-flow rates. ..........................................................................................68
xi

NOMENCLATURE

= spectral absorption coefficient [1/m]


= spectral intensity [W/(mm2sr)]
= spectral black-body intensity [W/(mm2sr)]
c = FLUENT cell edge length [m]
= specific heat capacity [J/(kg K)]
= number of photon emissions [ND]
G = irradiation [W/m2]
= thermal conductivity [W/(mK)]
L = receiver length [m]
m = mass [g]
= pressure [Pa]
= radiative heat flux vector [W/m2]
r = radius [m]
= time [s]
= velocity vector [m/s]
T = temperature [K]
u+ = wall coordinates velocity [ND]
u* = friction velocity [m/s]
= average velocity [m/s]
y = position [m]
y+ = wall coordinates position [ND]
z = axial position [m]
= under-relaxation factor [ND]
= viscosity [Pas]
= wavelength [m]
= standard deviation [ND]
= spectral scattering coefficient [1/m]
= spectral phase function [ND]
= non-dimensional emissive power [ND]
= wall shear stress [Pa]
= solid angle [sr]
= density [g/m3]
xii

ACKNOWLEDGEMENTS

I would like to acknowledge first and foremost my advisor for this research, Dr.
Fletcher Miller, whose guidance and experience made this otherwise-massive undertaking
possible, and whose enthusiastic belief in the potential of solar thermal power drew me to
this project in the first place. I would also like to acknowledge my predecessor in this work
Steve Ruther, who developed the original version of the Monte Carlo ray-trace radiation
solver for this receiver. My research would not have been possible without his work to build
upon. Dr. Arlon Hunt, who first proposed the concept of a small particle solar receiver, also
provided valuable insight throughout the progress of my research. The support and aid of my
colleagues in the SDSU Solar Energy and Combustion Laboratory was invaluable,
particularly Pablo del Campo who helped discover the final errors in the Monte Carlo code.
The financial support I received via grants from Google (grant # 32-2008) and the US
Department of Energy (grant # DE-EE0005800) not only made this work possible, but
allowed me to focus on this project full time.
1

CHAPTER 1

INTRODUCTION

Solar power has long been associated with renewable energy hopes, but has seen
significant growth in recent years. While photovoltaics have typically been the most high-
profile collection method, solar thermal power represents the best long term source for
utility-scale solar energy. Central receiver systems, sometimes referred to as power tower
systems due to their prominent collection towers, are one of the major configurations
available today for solar thermal energy production. In this arrangement, sunlight is focused
by a field of mirror clusters called heliostats onto a single receiver atop a tower as illustrated
in Figure 1.1 [1].

Receiver Sun

Generator Turbine Tower Heliostat Field

Figure 1.1. Schematic of a central receiver solar thermal power plant.


Source: Ruther, Steven James. Radiation Heat Transfer Simulation of a
Small Particle Solar Receiver Using the Monte Carlo Method. Masters
Thesis, San Diego State University, 2010.

The first large scale central receiver system to be built and tested was the Solar One
project in the Mojave Desert of California. Operating from 1982-1986, that system used
1,818 heliostats spread over 126 acres to boil water which drove a Rankine steam cycle,
producing up to 10 MW of electrical power [2]. It was redesigned and retrofitted in 1995 to
2

use molten salt as the receiver heat transfer fluid instead of steam, in the process becoming
Solar Two. The salt could then be stored in tanks before being sent through a heat exchanger
to boil water and drive the steam turbine. The concepts and technology demonstrated by
these two test projects are in use today in half a dozen commercial power plants in Europe
and the United States, as well as several currently under construction. The largest capacity
central receiver system in operation is the 20 MW Gemasolar plant, designed and operated
by Torresol Energy in southern Spain [3] and pictured in Figure 1.2 [4]. This power plant
began operation in 2011, and makes use of molten salt in the receiver and for storage in a
similar manner to Solar Two.

Figure 1.2. Aerial view of the Gemasolar power plant in Spain. Image
reproduced from Torresol Energy under the terms of the Free Art License.
Source: Torresol Energy. Gemasolar. Gemasolar.jpg. Last modified June 21,
2011. http://en.wikipedia.org/wiki/File:Gemasolar.jpg.

Solar thermal power takes advantage of over 100 years of thermodynamics research
and engineering by generating electricity from heat. This means that advances in turbine
design from existing fossil fuel research benefits solar thermal power as well. All existing
solar power plants built to date have made use of the Rankine steam cycle to produce
3

electricity. A temperature-entropy diagram of the Rankine steam cycle is shown in Figure 1.3
[5]. This is a phase-change cycle, which utilizes water and is commonly used at input
temperatures between 400-600 C. the Gemasolar plant for example has a turbine inlet
temperature of 565 C. This presents two major limitations for solar power. Being a phase
change cycle, cooling water is often needed to re-condense the working steam after going
through the turbine. For a technology best suited to dry desert climates, water usage is a
significant concern. Air cooling can and is used in some plants, but this comes with a
reduction in efficiency.

Figure 1.3. Rankine cycle diagram and schematic. Source: Stine,


William B. and Michael Geyer. Power Cycles for Electricity
Generation. Power from the Sun. Last modified October 2004.
http://www.powerfromthesun.net/Book/chapter12/chapter12.html#12.
4 Brayton Cycle Engines.

Secondly, a peak cycle temperature of 600 C is relatively low compared to modern


fossil fuel power plants, and thus imposes a lower thermodynamic efficiency limit on these
systems.
Both limitations of the Rankine cycle can be overcome by the Brayton cycle. This is
an all gas phase cycle, used in jet engines and some fossil fuel power plants. The Brayton
cycle can operate at much higher temperatures, and thus reach greater thermodynamic
efficiencies as compared to Rankine systems. Additionally, there is no need for cooling water
in an open, all gas phase cycle. A temperature-entropy diagram of an idealized Brayton cycle
is pictured in Figure 1.4 [5].
4

Figure 1.4. Brayton cycle diagram and schematic. Source: Stine,


William B. and Michael Geyer. Power Cycles for Electricity
Generation. Power from the Sun. Last modified October 2004.
http://www.powerfromthesun.net/Book/chapter12/chapter12.html#12.
4 Brayton Cycle Engines.

An important development in energy production has been combined cycle power


plants, which combine the efforts of Brayton gas turbine with Rankine steam generators to
produce higher efficiencies than either cycle individually. These systems work by using the
exhaust heat from a Brayton cycle to drive a Rankine generator, yielding a larger overall
temperature drop. The state of the art fossil fuel combined cycle generators are capable of
thermodynamic efficiencies as high as 60% today [6]. A diagram of the combined Brayton-
Rankine cycle is shown in Figure 1.5.
One challenge currently preventing the application of Brayton and combined
Brayton-Rankine cycles in solar thermal power is the ability to produce the necessary high
temperature gas. Until recently, the receivers in solar thermal systems have relied on surface
absorption of solar radiation before transferring the thermal input to a working fluid. A
traditional receiver design as used in Solar One is shown in Figure 1.6 [2]. In this
arrangement, incident sunlight from the heliostat field strikes the exterior of one of the many
dark vertical tubes which make up the circumference of the receiver body. The energy is
absorbed by the tube surface there, and is conducted through the tube wall to the interior,
where the cooling fluid carries it away via convection. In current plants, this coolant is
5

Figure 1.5. Combined Brayton-Rankine cycle diagram and schematic. Source:


Stine, William B. and Michael Geyer. Power Cycles for Electricity
Generation. Power from the Sun. Last modified October 2004.
http://www.powerfromthesun.net/ Book/chapter12/chapter12.html#12.4
Brayton Cycle Engines.

typically steam or molten salt. A major limitation of this design is the fact that the component
which absorbs incident radiation is not the working fluid itself, but rather an intermediary
tube. Thus, potential temperatures are limited by the material constraints of the absorber.
Additionally, the hottest part of the receiver is the exterior surface of the absorber tubes,
which are exposed directly to the environment. This can result in significant convection and
radiation losses, reducing the receivers efficiency. Current designs use absorber materials
with spectrally variant emission properties in an effort to reduce radiation losses, but these
materials can also have temperature limits. Limiting convection losses can be quite difficult
for traditional receiver designs.
The German Aerospace Center (DLR) has conducted research on a volumetric air
cooled receiver for the purposes of driving a gas turbine or combined cycle [7, 8]. The DLR
volumetric receiver operates at pressures up to 15 bar, and uses a porous absorber behind a
curved quartz window to collect concentrated solar energy. Air is forced through the absorber
and is heated to outlet temperatures of 800-1000 C. A diagram of the DLR receiver is shown
in Figure 1.7 [7]. This design was used and tested in conjunction with other lower cost and
lower temperature receivers. The high temperature volumetric receiver was used to heat
6

Figure 1.6. Diagram of the Solar One


receiver. Source: Stine, William B., and
Michael Geyer. Central Receiver Systems.
Power from the Sun. Last modified October
2012. http://www.powerfromthesun.net/
Book/chapter10/ chapter10.html.

pre-heated air from the other receivers to the target temperature of 1000 C, and the whole
system was tested in a hybrid scheme with a traditional fossil fuel gas turbine. Receiver
efficiencies averaged 80%, and pressure drop through the receiver was only 120 mbar.
Another potential receiver for high temperatures is the falling particle receiver
investigated by Sandia National Labs [9]. In this arrangement, small carbon particles
approximately 500 m in diameter are dropped through a beam of concentrated solar
radiation. A schematic of the receiver is shown in Figure 1.8. The particles can then be stored
directly or used to pass heat to a working fluid. At this size, the particles do not exhibit
selective absorption properties, and thus radiation losses can be significant. The falling
particle receiver has shown the potential to produce outlet particle temperatures in excess of
1000 C. Higher temperatures come at the cost of reduced efficiencies however. Another
limitation of this method is the difficulty in extracting the heat from the particles efficiently.
7

Figure 1.7. Schematic of pressurized high-temperature DLR receiver.


Source: Buck, R., E. Lpfert, and F. Tllez. Receiver for Solar-Hybrid Gas
Turbine and CC Systems (REFOS). Paper presented at the IEA Solar
Thermal Conference, Sydney, Australia, 2000.

Figure 1.8. Schematic of the Sandia falling particle receiver.


8

A new receiver design is being investigated here, in which the solar input is absorbed
volumetrically by a gas-particle mixture within the receiver cavity, thus heating the working
fluid directly. The advantages of this method include the ability to accommodate very high
incident flux levels due to volumetric absorption, reduced pressure loss, higher receiver
efficiency, and consequently higher outlet gas temperature. Higher outlet temperatures in
turn allow the possibility of utilizing a Brayton cycle for higher thermodynamic efficiency of
the overall system, ease of daily start-up and shut-down, as well as reduced water usage for
cycle cooling. A schematic of the new receiver design is shown in Figure 1.9.

Figure 1.9. Schematic of cylindrical small particle


solar receiver.

Small particle receivers are not limited, however, to electricity generation; the high
flux levels and intimate mixing between gas and particles offer many possibilities for solar
chemistry as well [10].
9

1.1 SPHER RECEIVER


A critical task in designing a small particle solar receiver is modeling the fluid flow
within the cavity, as well as the radiation heat transfer. These two pieces must work together
to give an accurate picture of receiver performance and behavior. The Small Particle Heat
Exchange Receiver (SPHER) was originally proposed independently in 1979 by both Hunt
[11] and Abdelrahman [12]. Its core features derive from the spectral variation in the
absorptivity of the small carbon particles. When controlled for size, the carbon particles
exhibit selective absorptivity, absorbing well in the solar spectrum and emitting poorly at
longer infrared wavelengths. With the proper mass loading of particles (mass of particles per
unit volume of air), the concentrated radiation can be absorbed in the volume of the receiver
before most of the radiation reaches a wall, thus keeping portions of the walls cooler than the
outlet fluid temperature reducing material constraints and heat losses.

1.2 MONTE CARLO RAY-TRACE METHOD


Due to the spectral and directional nature of the radiation source and the particle
properties, calculation of the radiation field in the receiver is difficult with traditional
methods. The Monte Carlo ray-trace (MCRT) method is a statistical approach to solving the
governing equation of radiation exchange in a participating medium, and can be used for
both radiative exchange between surfaces and participating media. It works by simulating
millions of discrete photon bundle emissions from every radiatively participating element in
the domain and tracing each path until the bundle is absorbed or leaves the domain.
Wavelengths, directions and locations of ray emissions are determined according to a
cumulative distribution function (CDF) computed according to the geometry and conditions
of the problem. Likewise, absorption, scattering, and reflection events are also determined
probabilistically based on the spectral and directional properties of surfaces and media in the
domain. Other methods of solving the RTE include the Discrete Ordinates method, which
solves the RTE analytically for a finite number of discrete direction vectors within the
medium. Radiation in directions other than those specifically solved for are lumped in with
the nearest solved direction. Because of this, the method has trouble with strongly directional
radiation. This method works reasonably well in optically thick mediums with uniform
directional radiation, but is less reliable near boundaries and in optically thin cases. The
10

MCRT method was selected for its ability to accurately model spectral property variations in
the system, as well as potentially complex input flux conditions on the receiver window from
the heliostat field. This input will eventually be taken from MIRVAL [13], which is another
Monte Carlo code. MIRVAL uses the MCRT method to simulate the behavior of a specific
the heliostat field, in this case the field at Sandia National Labs where the prototype receiver
will be tested.

1.3 SCOPE OF THESIS


This thesis models the radiation, convection, and conduction heat transfer along with
gas flow in an axisymmetric small particle solar receiver. This research expands on previous
work by coupling the MCRT radiation model developed by Ruther with the more
sophisticated fluid dynamics modeling capabilities of ANSYS FLUENT. This coupling
allows for the inclusion of more realistic inlet and outlet geometries in the receiver, similar to
those represented in Figure 1.9, which will be a critical part of the design of a real prototype.
The use of FLUENT also allows for a turbulence model, convection between the walls and
the gas, and much finer grid sizing for the energy solver. Taken together, these features give
a more complete picture of the heat transfer, mixing, and temperature profiles within the
receiver. The flow and energy solver remains two-dimensional and axisymmetric to save
computation time, while the Monte Carlo ray-trace is conducted in three dimensions. The
window is treated as an open aperture in the MCRT, and as a constant-temperature wall
boundary in the FLUENT solver. Other researchers in the group are working on optical and
cooling models of the window, and their work will be integrated with this model in the
future.
Background on previous experimental and modeling work on small particle receivers
is covered in Chapter two. A more detailed explanation of the Monte Carlo model is shown
in Chapter three, along with modifications to the MCRT model for this work which include a
Gaussian flux distribution on the aperture to more closely match flux maps produced by real
heliostat fields. Chapter three also touches on additional geometry changes to the MCRT
code which did not pass benchmarks and were thus not included in the final model for this
thesis, but have since been fixed and will be used in the future. An unheated flow through the
geometry is modeled in Chapter four in FLUENT, and the velocity profile is validated
11

against theoretical benchmarks. Computational and iteration techniques for coupling the two
solvers are investigated in Chapter five, including enforcement of minimum and maximum
temperature and source term values as well as several under-relaxation schemes. Receiver
performance is modeled at steady state conditions in Chapter six for a variety of mass flow
rates and inlet power levels. Gas outlet and cavity wall temperatures are computed and
compared with those obtained by Ruthers model, and receiver efficiencies are estimated.
12

CHAPTER 2

BACKGROUND ON SMALL PARTICLE


RECEIVERS

Experimental work on the small particle receiver concept was first done by Hunt and
Brown in 1982 [14]. That work demonstrated the potential of the design, producing outlet gas
temperatures of 1000 K, with a particle mass loading of 1 g/m3 and an incident flux level of
500 kW/m2. A schematic of the receiver built and tested by Hunt and Brown is shown in
Figure 2.1. The aperture window is at the bottom of the receiver as pictured here. The gas-
particle mixture entered the receiver through holes in a manifold on the cylinder wall, visible
in the middle of the diagram. Hot air exited an axial outlet tube upward, leaving the receiver
at the top.

Figure 2.1. Cross section of the experimental SPHER.

More recently in 2003, experiments by Bertocchi et al at the Weizmann Institute of


Science using sub-micron sized carbon particles and a particle mass loading as high as 7 g/m3
produced outlet gas temperatures of 2000 K. [15]. The receiver in those tests was
significantly smaller than that investigated here, with an aperture diameter up to 80 mm and
13

average incident flux levels up to 900 kW/m2, though peak concentration reached 5 MW/m2.
The flow orientation also differed from present work, as the Weizmann receiver had inlets
near the aperture with the bulk of the gas-particle mixture flowing concurrent with the
incident radiation. This flow orientation is illustrated in Figure 2.2 [15] which shows a cross-
sectional schematic of the Weizmann receiver.

Figure 2.2. Schematic Cross section of the Weizmann


Institute receiver. Source: Bertocchi, R., J. Karni, and
A. Kribus. Experimental Evaluation of a Non-
Isothermal High Temperature Solar Particle Receiver.
Energy 29, no. 5-6 (2004): 687-700.

That work investigated variations in mass flow rate, particle mass loading, and
aperture size, as well as exploring several gases. Receiver efficiencies were estimated
between 80-90%. Further research by Klein, Rubin, and Karni [16] at the Weizmann Institute
in 2008 produced exit gas temperatures up to 1500 K under flux levels of 3 MW/m2. That
work also demonstrated wall temperatures below peak gas temperatures, a finding first seen
in the original tests conducted by Hunt and Brown [14].
Early numerical modeling of small particle receivers was undertaken by Miller in
1988 [17]. That work modeled an experimental lab scale cylindrical receiver 1 m long and
14

5.9 cm in diameter, powered by a xenon arc lamp with a mostly collimated input. Flow could
be oriented concurrent or opposed to the incident radiation. The model used a four-flux
method to solve the radiative transfer equation in the gas-particle mixture. While specular
variations were considered, the primary limitation of this method was an inability to
accurately capture the effects of scattering and wall boundaries. Additionally, the gas flow
was modeled as either a laminar parabolic profile or a flat slug flow. Temperatures in the gas
flow of the experiment proved difficult to accurately measure, and the model did not produce
good agreement with experimental results.
Numerical modeling of the Weizmann receiver was done by Klein et al. in 2006 [18],
utilizing commercial CFD software coupled with a Monte Carlo ray-tracing method.
However, that work used the Monte Carlo method only for surface radiation exchange, and
used a discrete ordinance method for radiation heat transfer within the gas-particle mixture.
Additionally, the Monte Carlo model was gray and thus unable to capture the spectral
absorption properties of the particle cloud. The flow model in that work was laminar, as the
geometry of the smaller receiver resulted in a low Reynolds number of 400. By contrast, the
Reynolds number in the main cavity of the receiver studied in this thesis is 4000. Good
agreement was found between experimentally observed and numerically computed wall
temperatures.
The Monte Carlo ray-trace method has also been used to model heat transfer in a
solar chemical reactor by ZGraggen and Steinfeld [10]. The solar reactor shares many
features with small particle solar receivers including a quartz window for incident radiation, a
cylindrical cavity, and direct irradiation of small particles. Unlike small particle receivers
however, in solar chemical reactors the concentrated sunlight is used to drive chemical
reactions instead of simply produce hot gases. Like the Weizmann experiments, the
ZGraggen reactor includes particle-gas flow concurrent with the incident solar radiation, a
smaller scale design than that investigated here (97mm reactor diameter), and peak flux
levels up to 5 MW/m2. That model also used a finite volume CFD solver coupled with a
Monte Carlo ray-trace code to simulate heat transfer in the reactor. In this case, the Monte
Carlo model included spectral considerations and was used for both surface radiation
exchange as well as in the participating medium. Solar input was used to raise temperatures
to 1300 K where chemistry took over and drove temperatures as high as 2000 K. Again,
15

modeling did a good job of predicting wall temperatures in the reactor, though gas
temperatures in the cavity were less easily matched by experiments.
A concern common to all these various implementations is the structural integrity of
the quartz window, required either to separate gases and reactants from the environment or to
contain pressure. Window investigations have been undertaken in modeling by Rger, Buck,
and Muller-Steinhagen at the German Aerospace Center [19] as well as experimentally by
Karni et al. at the Weizmann Institute [20]. Both experiments used small windows from 12-
32cm in diameter, and found it feasible to maintain a quartz window under pressure up to 15
bar and with internal receiver temperatures in excess of 1300K. Modeling of a larger window
for use in this project was done by Mande [21], though that work was limited to a stress
analysis and did not include temperature considerations. The quartz window is a pivotal but
complex piece to this receiver design, and as such is outside the scope of this thesis. It is a
subject under active research by others in the group at San Diego State University.
16

CHAPTER 3

PREVIOUS WORK AND MODIFICATIONS TO


MONTE CARLO RAY-TRACE MODEL

Previous work on the current receiver has been done by Steve Ruther, using the
Monte Carlo Ray Trace (MCRT) method to model the detailed radiation heat transfer within
a cylindrical receiver with absorbing, emitting, and scattering particles [1]. As introduced in
Chapter 1, the Monte Carlo method is a statistical strategy for solving the radiative transfer
equation (RTE) which is the governing equation of radiation exchange in a participating
medium and is shown in Equation 3.1 [22].

(3.1)

The radiative transfer equation characterizes the change in spectral intensity, , as radiation
travels along a path length through the medium. It includes four terms: energy lost to the
medium via absorption, energy gained via emission from the medium, energy lost due to
scattering, and energy gained due to in-scattering. Each of these terms can have spectral
dependence, and the scattering terms can be highly directional as well. The RTE must be
integrated over a path length to determine the value of spectral intensity at any point in the
domain. That intensity can then be used in an energy balance as shown in Equation 3.2:

(3.2)

This is an energy balance between total energy gained and lost via radiation in a control
volume. The sum is expressed as which is the divergence of the radiative flux.
In order to obtain temperature profiles in the gas-particle mixture, the RTE must be
coupled to the conservation of energy equation. This is accomplished by including the
divergence of the radiative flux in the energy equation of the gas flow as a source term, as
illustrated in Equation 3.3. In this formulation, the divergence of the radiative flux represents
net energy added or removed from a control volume via radiation.
17

(3.3)

The full conservation of energy equation includes the total derivative of the
temperature, which in turn includes a velocity term due to advection. Thus it is necessary to
solve for the flow field as well in the model to determine the temperature field. Ruthers
work included a uni-directional slug flow model of the gas particle mixture in order to obtain
velocity profiles for the energy equation. In that model, the energy equation was simplified to
include only axial advection, radial diffusion, and the volumetric source term for radiation.
Both the Monte Carlo ray-trace and energy solver portions of the model were written in
FORTRAN. The geometry was limited to orthogonal boundaries in order to simplify the
Monte Carlo code and keep computation times relatively short. The window was set at 1.5 m
radius to be at what was considered the upper limit for a quartz window. The total receiver
radius was set at 2.5 m to investigate heating of the gas from indirect insolation (i.e.,
scattered light), and the axial length was set at 5 m to allow for a wide range of optical
thicknesses in the participating medium.
Due to the simplified nature of the energy solver and flow model, Ruthers work
focused on the radiation heat transfer. Variables studied included particle mass loading,
particle size, mass flow rate, wall optical properties, and flow direction. The greatest receiver
efficiencies of 90% were predicted with opposed flow, entering at the rear of the receiver
cavity and traveling forwards towards the incident radiation, with a particle diameter of 0.2
m and particle mass loading of 0.30 g/m3. Peak temperatures of 1600 K were predicted,
with a 5MW solar input and a uniform flux level of 707 kW/m2. Figure 3.1 shows an
example of the output from Ruthers version of the model. It shows the axisymmetric
temperature field in the gas-particle mixture for a 5MW collimated input, with 0.30 g/m3
particle mass loading and 0.5 m particle diameter.

3.1 INCIDENT FLUX


Previous work included two different arrangements of incoming solar radiation
entering the receiver: uniform collimated radiation as a base case, and a 45-degree cone angle
to more accurately represent what comes from the heliostat field [1]. Both schemes assumed
a uniform incident flux on the aperture, varying only the angle of incidence. This uniform
flux assumption is however not realistic, as real world heliostat fields cannot reasonably
18

Axisymmetric Temperature Field


Color Bar Units K
1600
3
1500
2.5
1400
2
1300
Raduis (m)

1.5 1200

1 1100

0.5 1000
900
0
800
-0.5
700
1 2 3 4
Length (m)
Figure 3.1. Temperature field output from original MCRT code. Flow is
from right to left, as indicated by white arrows.

provide a uniform flux over the course of a day, and in fact rarely produce a uniform flux
map at all. This is illustrated in Figure 3.2, which shows an example flux map computed by
the MIRVAL code, which is a Monte Carlo simulation of the heliostat field at Sandia
National Labs [13].
In the current absence of a full heliostat field model coupled to the receiver, the
receiver model has therefore been modified to apply a Gaussian flux distribution to the
aperture. The original model computed the bundle energy of each emission from the aperture
according to Equation 3.4 where is the bundle energy, dA is the area of the individual
aperture element, and flux is the incident solar flux on the aperture element.

(3.4)
#

By rearranging this formulation, the number of emissions can be varied based on a bundle
energy consistent with the original model, and dependent on the desired flux on the aperture
element. The aperture plane has five radial divisions, to create five concentric ring-shaped
elements. The uniform flux case imparted 5 MW on a 1.5 m radius aperture, resulting in
19

Figure 3.2. Flux map predicted from MIRVAL. Figure


represents 3 m by 3 m region on aperture plane. Source:
Leary, P. L., and J. D. Hankins. A User's Guide for MIRVAL
- A Computer Code for Comparing Designs of Heliostat-
Receiver Optics for Central Receiver Solar Power Plants.
Livermore, CA: Sandia Labs., 1979.

0.707 MW/m2. For the Gaussian distribution, the peak flux is chosen to be just short of thrice
this, at 2.8 MW/m2. This peak was chosen based on simulations of the heliostat test field at
Sandia National Labs using the MIRVAL code. For a Gaussian flux distribution centered at
the middle of the aperture, the probability weighting for a given location is given by equation
3.5 where is the standard deviation, and r is the radius from the receiver centerline.

(3.5)

The local flux at each radial division is weighted by this probability, and a standard deviation
of 0.52 yields a total power input equal to the uniform flux case of 5 MW. The flux
distribution on each aperture ring element is plotted in Figure 3.3, along with the MIRVAL
20

Gaussian Flux Distribution


3.0
Approximated Model Flux
2.7
Calculated Real Flux

2.4
Local Incident Flux [MW/m2]

2.1

1.8

1.5

1.2

0.9

0.6

0.3

0.0
0.15 0.45 0.75 1.05 1.35
Radius of Aperture Element [m]

Figure 3.3. Plot of the local incident solar flux applied to the aperture.

results for comparison. The wavelength distribution of the incoming radiation is modeled as
blackbody emission at 5780 K to approximate the solar spectrum.

3.2 OUTLET TUBE


The major modification attempted to the original Monte Carlo model is the addition
of an outlet tube to the geometry. The outlet tube exists as a smaller cylinder concentric with
the outer cylindrical wall. It is intended to absorb, emit, and reflect rays in the same manner
as the other receiver walls. To add the outlet tube to the model, an additional radial division
is needed to represent tube wall elements. This radial division has zero thickness, like all wall
elements in the domain. A schematic of the modified geometry with the outlet tube is shown
in Figure 3.4, where r1 is the radius of the outlet tube, r2 is the radius of the outer cylinder
21

Figure 3.4. Schematic of modified receiver geometry in Monte


Carlo ray trace model.

wall, and g is the gap distance between the entrance end of the outlet tube and the aperture
plane of the receiver.

3.3 COMPARISON TO BENCHMARKS


To verify the outlet tube addition to the Monte Carlo model, a comparison is made
with existing solutions. The first two benchmark cases investigated examine radiation
exchange between surfaces, with no participating medium between them. The first of these is
meant to model radiation exchange between parallel infinite plates. To simulate this in the
Monte Carlo code, the geometry is modified by extending the outlet tube to span the full
length of the receiver axis. This change will carry over to all the benchmarks described in
this chapter. For the parallel plate case, the end walls are treated as black bodies at 500 K,
while the cylinder walls are perfect specular reflectors as illustrated in Figure 3.5. Results
from this case are plotted in Figure 3.6, with radial position oriented on the vertical axis.
The plot in Figure 3.6 shows the net energy gained or lost by each ring element on
either end wall. The results show a clear pattern, where the net source terms are negative and
of similar magnitude for all end-wall ring elements except those adjacent to a cylindrical
wall, which are significantly greater in magnitude. Although the energy balances between the
22

Figure 3.5. Schematic of modified receiver geometry in Monte


Carlo ray trace model for black end-walls benchmark.

Energy Exchange Between Black End-Walls


5.0

4.5

4.0

3.5
Radial Distance from Axis [m]

leftwall
3.0
rightwall
2.5

2.0

1.5

1.0

0.5

0.0
10,000 5,000 0 5,000 10,000 15,000
Net Energy Gain in Wall Element [W]
Figure 3.6. Net energy exchange between parallel black plates with no participating
medium, both at 500 K. The thick black line at 2.0 m represents the outlet tube.
23

two walls, three elements on each wall show net energy changes significantly removed from
zero. Since both walls are set at a uniform 500 K and the system is completely insulated, the
system is already at equilibrium and thus the net radiation exchange at every wall element
should be exactly zero according to the analytical solution. Because the Monte Carlo method
is a statistical simulation of radiation heat transfer, some deviation from zero is to be
expected. But for the benchmark case to match expectations, these deviations from zero
should be small in magnitude and randomly distributed throughout the domain. The results in
Figure 3.6 show a clear bias, indicating a problem with ray tracing in this case.
The second benchmark case without a participating medium is similar to the first. The
only modification is to the emissive properties of the walls. In this case, the end walls are
modeled as perfect specular reflectors, and the concentric cylinder walls are modeled as
black bodies with an emissivity of one. This is illustrated in Figure 3.7. The inner cylinder is
set at a constant 500 K, while the outer cylinder is set at a constant 1000 K. With the end
walls being mirrored, the geometry should model infinite concentric cylinders.

Figure 3.7. Geometry of Monte Carlo model for concentric


cylinders benchmark.

Results from this case are plotted in Figure 3.8, which shows the net energy gained by
each concentric wall element according to axial position. Because of symmetry, there should
not be any axial change in reported energy exchange for this case, but a clear trend is visible
in the net energy gained by the inner cylinder. This again indicates an error in the ray tracing
with the inclusion of the outlet tube.
24

Energy Exchange Between Concentric Black Cylinders


1.2
InnerCylinder(500K)
1.0
OutterCylinder(1000K)

0.8
Net Energy Gain in Wall Element [MW]

0.6

0.4

0.2

0.0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
0.2

0.4

0.6

0.8
Axial Location Along Wall [m]

Figure 3.8. Net energy exchange between concentric cylinders with no participating
medium. The inner cylinder is at 500 K and the outer cylinder is at 1000 K.

The next benchmark case to consider is radiation in a gray absorbing-emitting medium


between infinite parallel plates [23]. This comparison was also utilized by Ruther [1] to
validate the first version of the MCRT model. In this case, the geometry is similar to the first
case presented in this section, only with the addition of a participating medium. The end
walls are modeled as black, while the cylinder walls are treated as perfect specular reflectors.
The left end wall is set to a constant 1000 K while the right end wall is set to a constant 500
K. Only radiation heat transfer is considered in the domain, and the gas is given a uniform
absorption coefficient. Theory predicts this should result in a linear axial non-dimensional
temperature profile in the gas. The geometry of this benchmark is identical to that illustrated
previously in Figure 3.5. For this case, the geometry is again modified to extend the outlet
tube along the entire axis of the geometry, such that it intersects both end walls. This,
combined with both cylindrical walls being specular with zero emissivity should make the
25

heat transfer match the benchmark case of infinite parallel plates. Results of model are
plotted in Figure 3.9, which shows non-dimensional gas temperature as a function of axial
distance from the hot wall for four different radial positions, all of which are between the two
concentric tube walls. This is compared to results from Howell and Perlmutter [23] for an
optical depth of 2. The temperatures in the gas are normalized according to equation 3.6.

(3.6)

The temperature profile is as expected at three of the radial positions, but not at the innermost
division of r = 0.45 m. This is the radial gas division adjacent to the outlet tube, and a
temperature anomaly here implies unexpected behavior on the outlet tube wall. Further
exploration of this issue is detailed in the Appendix.

Emissive Power in Gray Plane Layer


1

0.9
Non-dimensional Temperature ( )

0.8

0.7

0.6

0.5

0.4 r=0.45
0.3 r=0.75
r=1.05
0.2
r=1.35
0.1
benchmark
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Non-Dimentional Axial Distance (z/L)

Figure 3.9. Plot of non-dimensional emissive power in gray gas plane layer benchmark.

The final benchmark appropriate to the new geometry is radiant transfer between
infinite black concentric cylinders bounding a gray absorbing-emitting medium [24]. This
allows for a test of absorption and emission on the central outlet tube wall, while keeping the
computation time down by excluding scattering. To run this case in the MCRT code, the
26

outlet tube length is again extended to both end walls and the end walls are modeled as
perfect specular reflectors as shown in Figure 3.7. The gas volume is divided into 20 radial
elements by 5 axial elements, in order to provide additional resolution in the radial direction.
This scheme provides improved resolution in the radial direction for comparison to the
benchmark, while reducing axial divisions to save on computation time. The geometry is
modified to match the benchmark cases in Perlmutter [24]. The inner radius (r1) is 0.15m and
the outer radius (r2) is 1.5m, yielding a cylinder radius ratio r1/r2 = 0.1. The cylinder walls are
given an emissivity of 1, and three optical thicknesses (0.1, 2, 10) are modeled. The optical
thickness is defined as = a(r2 r1) where a is the absorption coefficient in the gas. This
results in an absorption coefficient for the three cases of 0.074 m-1, 1.481 m-1, and 7.407 m-1.
The results are plotted as non-dimensional location vs. emissive power. The radial location
(r) in the gas is non-dimensionalized as follows.

(3.7)

The emissive power is again non-dimensionalized by Equation 3.8, where T1 and T2 are the
temperatures in Kelvin of the inner and outer cylinders, 1000K and 500K respectively.

(3.8)

Figure 3.10 shows the radial profile of emissive power in the gas for an optical thickness of
10, at each of the axial divisions in the gas. The results show no variation in the axial
direction, indicating that the specular end wall properties provide a sufficient substitute for
infinite cylinder length.
The results for each of the three optical thicknesses considered are shown in Figure
3.11, denoted by in the plot legend. For these data, the emissive power at each radial
division was averaged across all axial divisions to provide a larger statistical basis. The
benchmark data is extracted from an image of the results plotted in Perlmutter [24], with the
aid of the Plot Digitizer program [25]. Agreement with the benchmark is reasonable for an
optical depth of 2, but greater and lesser optical thicknesses yield poor results when
compared to the respective benchmark cases. There appears to be little difference in the
model results between 0.1 and 2 optical thickness cases, while the largest optical depth case
27

0.3
EmissivePowerinGrayGas = 10
0.25
z=0.5m
NonDimensionalEmissivePower

z=1.5m
0.2 z=2.5m
z=3.5m
z=4.5m
0.15

0.1

0.05

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
NonDimentionalDistanceAcrossAnnulus

Figure 3.10. Plot of non-dimensional emissive power in gray gas between concentric
cylinders. Each data set represents the radial temperature profile at a different axial
location.

of 10 indicates a trend in the wrong direction. The appendix catalogues more detailed
investigation of these issues, including cases without any participating medium as well as
one-wall emissions and extreme aspect ratios in an attempt to isolate the problem.
Unfortunately, despite considerable effort, the underlying source of these errors was not
found in time for corrected results to be included in thesis1. As a result, the coupled model
described in the remainder of this thesis uses the original version of the MCRT code, without
an outlet tube. In other words, the outlet tube is included in the fluid dynamic calculations,
but is considered transparent to radiation in the MCRT portion of the code. A single
comparison with and without the outlet tube, explained later, showed little difference in

1
As of final publication of this thesis the error was found and corrected and good agreement with the
benchmarks was obtained. The MCRT with outlet tube will be used for future calculations.
28

0.8
EmissivePowerinGrayGas Concentric
0.7 Cylinders
tau=0.1(Benchmark)
NonDimensionalEmissivePower

0.6 tau=0.1(MonteCarloResults)
tau=2(Benchmark)
tau=2(MonteCarloResults)
0.5 tau=10(Benchmark)
tau=10(MonteCarloResults)
0.4

0.3

0.2

0.1

0.0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
NonDimensionalDistanceAcrossAnnulus(r/R2)
Figure 3.11. Plot of non-dimensional emissive power in gray gas between
concentric cylinders for three different optical depths.

receiver outlet temperature or efficiency. The primary benefit of having the outlet tube in the
radiation model is to determine its temperature (and hence material) for constructing a real
receiver.
29

CHAPTER 4

COUPLING OF THE MONTE CARLO AND


FLUENT MODELS

This chapter describes the methods by which the modified Monte Carlo and FLUENT
Computational Fluid Dynamics models are coupled to solve for the temperature field in the
receiver. A User Defined Function (UDF) was written for this thesis in C and is used in
FLUENT to call custom code during each iteration of the flow solver. This enables the
Monte Carlo code to compute the radiation heat transfer, while FLUENT solves the
conservation equations of mass, momentum, and energy. The User Defined Function (UDF)
utilizes four DEFINE macros in FLUENT: DEFINE_EXECUTE_ON_LOADING,
DEFINE_SOURCE, DEFINE_PROFILE, and DEFINE_EXECUTE_AT_END. The first is
called when the UDF is loaded into FLUENT, and reads initial source term data from a data
file. The second segment is called each time FLUENT solves the energy equation for a
volume cell in the domain. It applies the appropriate source term to the cell based on the
cells location. The third subroutine applies source terms to the outer cylinder wall. The final
UDF is run at the end of each Fluent iteration. It queries the temperature field from FLUENT
and puts this data into an array, then passes the array to the FORTRAN code, which executes
the Monte Carlo iteration. An additional subroutine will need to be written to apply source
terms to the outlet tube wall in FLUENT, now that the MCRT code with an outlet tube is
validated. Because the outlet tube has radial thickness in FLUENT, unlike the outer cylinder
wall, it will require its own UDF.
The coupling scheme is illustrated in Figure 4.1. The process begins with the
initialized data in FLUENT. This data consists of an initial flow and temperature field
acquired typically from a previous case. Additionally, the initialized data includes initial
source term estimates loaded from a data file, or computed from a previous case. For new
cases at very different inlet power and mass-flow rates, the initial flow field is acquired by
running the FLUENT solver uncoupled from the MCRT at the target mass flow rate with no
source terms applied and a constant 1000 K gas temperature. The program solves the
30

Figure 4.1. Simple flow diagram of the coupling scheme used by the model.

conservation of energy, momentum, and mass equations to produce temperature, pressure,


and velocity fields in the domain. The temperature field is passed to the Monte Carlo ray
trace program, which uses it to compute the radiation heat transfer. The MCRT computations
produce an array of net volumetric and wall source terms which are passed back to FLUENT.
The flow solver then applies these source terms, re-computes the temperature field (and new
velocity and pressure fields as well), and the two pieces continue to iterate until convergence.
The model is run on a Dell PC running the Windows 7 operating system, with two 64-bit
quad-core Intel Xenon X5560 processors running at 2.80 GHz and 6 GB of RAM. Because
the Monte Carlo code is not set up for parallel computations, each case runs only on a single
processing core, though multiple cases are run on the machine at once.

4.1 COUPLING SCHEME & UDF FUNCTIONS


DEFINE_EXECUTE_ON_LOADING:
This function is run only once for each simulation, when the volumetric and wall
source term data is first loaded into FLUENT. It initializes the iteration counter,
loads preliminary volume and wall source terms from a data file, and sets the
number of Monte Carlo ray emissions to 10,000 per cell for both volume and wall
surface elements.
DEFINE_SOURCE:
This function is called once for each volume cell in the FLUENT domain on
every iteration. When FLUENT solves the energy equation for a volume cell, this
function is called and the cell information is passed to it. The built in FLUENT
command C_CENTROID is used to query the centroid location of the cell being
solved. Two nested loops and an IF statement then check which MCRT cell the
current FLUENT cell is within, based on its centroid location. This set of IF
statement checks is hard-coded for the uniform MCRT grid used for this thesis. A
non-uniform Monte Carlo grid would require a revision to this function. The
31

corresponding volumetric source term is then applied to the FLUENT volume cell
to end the function.
DEFINE_PROFILE:
This function is called once per FLUENT iteration, with the purpose of applying
source terms based on the radiation model to the outer cylinder wall in FLUENT.
To accomplish this task, the function has two nested loops and a series of IF
statements. The outer loop steps through every face element in the FLUENT
domain along the outer cylinder wall. Each step, the centroid of that face is
queried using the built-in FLUENT function F_CENTROID and its axial location
is checked against the MCRT grid boundaries. Since the wall source terms passed
from the MCRT solver are a step function with divisions every 0.5 m, linear
interpolation is used to apply a continuous flux function to the cylinder wall in
FLUENT. Figure 4.2 shows an example of the raw wall flux values from the
MCRT (in blue) plotted against the interpolated wall flux values applied in
FLUENT (in red).
Equation 4.1 shows how this interpolation is computed within the subroutine,
where & are the reference flux values from the MCRT solver on either side
of the FLUENT face cell under scrutiny, z is the axial coordinate of the face
centroid, and z1 is the axial location of the reference MCRT element.
2 (4.1)
The reference flux values used in this interpolation are computed using an under-
relaxation factor detailed in the next section.
DEFINE_EXECUTE_AT_END
This function is the longest of the four UDF subroutines, and accomplishes
several tasks. It is executed once at the end of each FLUENT iteration. The first
thing this function does is compile the temperature field data and save it in an
array that the MCRT solver can use. This is done with three nested loops; the
outer two step through each of the 50 MCRT volume cells and the inner most
loop steps through every FLUENT volume cell in the gas domain. Within the
inner loop, each cell centroid is queried, again using the built-in FLUENT
function C_CENTROID. The cell coordinates are checked against the boundaries
of the current MCRT cell being looped through. If the FLUENT cell centroid is
determined to lie within the present Monte Carlo cell, the built-in FLUENT
function C_T is used to query the FLUENT cells temperature. The fluent cells
mass is then computed based on the cell volume and density, both of which are
given again by FLUENT functions. Finally, this temperature is added to the
running total temperature of the present MCRT cell, weighted by the mass.
Equations 4.2 & 4.3 show how the running totals of mass and temperature of each
MCRT cell are computed, where and are the temperature and mass of an
individual FLUENT cell, respectively. Likewise is the running total mass
of an individual MCRT cell which contains the given FLUENT cell. A running
32

Figure 4.2. Wall Flux Interpolation Method. Vertical grid lines correspond to wall
element divisions in the Monte Carlo solver. A negative net wall flux means that the wall
cell is a net emitter of radiation and thus is heated by the fluid, while a positive value
means the cell is a net absorber of radiation and is cooled by the gas.

total sum of the product of each FLUENT cell temperature and mass, represented
by , is computed in Equation 4.3.
, , (4.2)
4.3)
At the end of each MCRT cell loop, when all FLUENT cells within it have been
queried and added to the running total, the final mass-weighted average
temperature of the MCRT cell is computed according to Equation 4.4.
(4.4)

The next task of this UDF subroutine is to aggregate the outer cylinder wall
temperatures using a similar method. For the wall elements, the average
temperature in each Monte Carlo division is computed based on the area of each
FLUENT face since they do not have mass or volume in the model. Again each
33

Monte Carlo division is stepped through, and within each one the subroutine
loops through every FLUENT face on the outer cylinder wall, adding the
temperatures to the running total of the appropriate MCRT element. At the end of
this process however, the temperature array intended for the Monte Carlo must be
checked for extremes. This is because the data tables generated by the cumulative
distribution functions of the MCRT solver must have finite temperature ranges set
in advance of model computations. These limits are set at 200 K and 2200 K, so
any average wall temperatures beyond these limits are set to the closest acceptable
value. This check was not necessary for the volume cells because FLUENT
allows enforcement of temperature limits on volume cells during its calculations,
so the UDF does not need to do it. After both wall and volume temperatures are
aggregated into an array that can be passed to and read by the MCRT solver, a
pair of nested loops steps through each element of this array and applies an under-
relaxation scheme to the temperature data, the details of which are discussed in
the following section.
The next task of this function is one of managing the iteration process. The
iteration counter is incremented, and the number of emissions from each element
in the following Monte Carlo ray trace is determined. The first 50 iterations use a
static 10,000 emissions per cell, to keep computation time down while the model
is still far from the converged solution. After the 50th iteration, the number of
emissions is increased by 10% each iteration until reaching a maximum of
1,000,000 per cell. This ramp up of emissions is similar to the method used by
Ruther in the first version of the model. Additionally, Ruther showed that 1
million emissions per cell are needed to reach acceptably accurate radiation heat
transfer calculations [1]. The present model iterates 5 times once the 1 million
emissions threshold has been reached, giving a total of 105 iterations.
The last step before this subroutine calls the FORTRAN MCRT solver is to
generate a seed value for the random number generator needed in the Monte Carlo
code. This is done by querying the system clock, and using the five-digit time in
seconds as the seed. The FORTRAN programing language lacks the built-in
functions needed to query the computer systems clock, but the UDF is written in
C which does have this feature. Thus, each iteration the Monte Carlo is given a
new seed value for its random number generator. This allows for the model to
iterate an indefinite number of times if need be, without the random number
sequence produced in the FORTRAN code repeating itself. The eventual
repetition of the random number sequence placed a practical limit on the number
of iterations Ruthers version of the model could do while remaining valid [1].
Finally, the UDF is ready to call the FORTRAN subroutine which holds the
Monte Carlo solver. The values passed to the MCRT code include the temperature
field, the number of emissions per cell, and the random number seed. The Monte
Carlo passes back to the UDF volumetric source terms and wall flux values. After
these values are computed and passed back to the UDF by the MCRT, they are
written to a text file for data keeping purposes. Additionally, the wall flux values
are checked for extremes, as described in more detail in the following section.
34

After this final UDF subroutine is complete, the FLUENT solver begins its next
iteration.

4.2 SOLUTION CONTROLS


Because of the coupled nature of the model, combined with the numerical variation
introduced by the Monte Carlo method, several solution controls are necessary to reach a
stable convergence. The first method introduced is clipping of source terms applied to the
outer receiver wall. Early on during model iteration when the total number of Monte Carlo
emissions from each element is relatively low and the temperature field is still developing,
the net wall source terms returned to FLUENT from the MCRT can be of extremely large
magnitude. This presents a problem, as the energy balance at the receiver wall is described
by Equation 4.5, where is the net radiative source term at the wall from the MCRT, in
W/m2.

(4.5)

The thermal conductivity of the gas is proportional to the gas temperature raised to
the 3/4ths power, but this is approximated in the model by a linear function of temperature.
The thermal conductivity is in the range of 0.05-0.15 Wm-1K-1 at temperatures found in the
receiver. This gives an upper limit to the magnitude of the heat flux at the wall, as the gas
temperature beyond the thermal boundary layer is driven by advection in the cavity and
volumetric absorption from incident radiation, and thus is not significantly influenced by
conduction from the wall. This can place significant constraints on the grid near the wall. To
illustrate this, consider Figure 4.3 which plots the gas temperature in the thermal boundary
layer near the outer cylinder wall at four different axial locations. The points on each contour
represent actual grid points in FLUENT.
The green line represents the boundary layer at 4.25m from the aperture, or 0.75m
from the gas inlet. At this location, the wall is heated to 870K by radiation exchange in the
cavity, but the gas is only a bit below 750K, near the inlet temperature of 700K. The Monte
Carlo solver produced the maximum flux value of 5kW/m2 for this wall cell ( in Equation
4.5). The thermal conductivity of the gas (k in Equation 4.5) in this temperature range is
approximately 0.06 W/(m K), and the temperature change through the beginning of the
boundary layer (the first two grid points) is 110 degrees Kelvin. To produce the required
35

1100

1050
z=0.25
z=0.75
1000
z=1.75
Gas Temperature [K]

z=4.25
950

900

850

800

750

700
0 2 4 6 8 10 12 14 16 18 20
Radial Distance from wall [mm]
Figure 4.3. Thermal boundary layers near the outer cylinder wall at four different axial
locations. Legend values indicate axial distance from aperture of each data series.

temperature gradient to balance Equation 4.5, the wall temperature of 870K and gas
temperature of 760K would need to be separated by only 1.3mm. But as Figure 4.3 shows,
the first gas cell adjacent to the wall is located roughly 2.5mm away. Thus, the grid is too
coarse to resolve the boundary layer in this region. Further refinement of the FLUENT grid
in this region may be desirable, but it would come at the cost of increased computational
load. As discussed in Chapter 5.3, while a finer grid is better capable of resolving these
boundary layers, it does not significantly change the outlet temperature and receiver
efficiency. Figure 4.3 also illustrates how the grid is in fact fine enough to resolve the
thermal boundary layers at other axial positions along the outer cylinder wall. Wall heat flux
values an order of magnitude or more than this result in unrealistic solutions from the
FLUENT solver, such as negative temperatures with large enough negative fluxes. Thus, the
incoming wall source terms from the Monte Carlo solver are checked each iteration and
limited to an absolute value of 5 kW/m2. This prevents physically impossible temperature
36

results due to the grid sizing, along with limiting large fluctuations that occur early on in the
iteration process.
The second solution control method is the introduction of an under-relaxation factor
for all source terms computed by the Monte Carlo solver. These are applied each iteration,
before the source terms are used by FLUENT to update the temperature field. This damps out
some of the inherently unstable nature of the solution, which is again particularly evident
early on in the iteration process. A relatively aggressive under-relaxation factor of 50% is
used on all source terms. This control method represents one half of the strategy to apply
under-relaxation to the solution of the energy equation, with the other half being applied to
temperature values.
The final solution control used on the model is an under-relaxation factor on the
temperature field. This is applied each iteration to the temperature field produced by
FLUENT, before it is passed to the MCRT. Equation 4.6 shows how the temperature of each
MCRT element is calculated, where Ti is the cell temperature computed by FLUENT during
the current iteration, Ti-1 is the cell temperature applied to the MCRT code on the previous
iteration, and is the under-relaxation factor.

1 (4.6)
It should be noted that this method is based on T4 and not simply T, because radiation is the
dominant heat transfer mode in the gas and is dependent on T4. The importance of solution
controls on T4 as opposed to T is illustrated in Figure 4.4, which plots the net energy gain by
the receiver (as computed in the MCRT code) for a single case under each scheme. For the
first 20 iterations of this case (in blue), under-relaxation was applied to T4, and from iteration
21 onward (in red) it was applied to simply T. It is clear that the former scheme is more
stable than the later, despite both being applied to a case that was already very near
convergence. Figure 4.5 shows the same case with no under-relaxation on the temperature at
all. Not only are the oscillations in this case much greater, but their magnitude actually grows
over the course of the model run, showing significant instability.

4.3 GRID MATCHING SCHEME


As is discussed in more detail in Chapter 5.1, the FLUENT grid divisions are
generated without concern for the locations of the Monte Carlo grid divisions. The coarsest
37

Effect of under-relaxation on T4 vs. T


4.80
Millions

4.75
Net Energy Gain in Reciever [W]

4.70

4.65

4.60

T^4
4.55 T

4.50
0 10 20 30 40 50 60

Iteration number
Figure 4.4. Net energy gain in the gas flow as a function of model iteration number for
two different temperature under-relaxation methods.

part of the grid has an average of 200 FLUENT cells in each MCRT cell, while the finest
region has several orders of magnitude more than this. To ensure that the FLUENT and
MCRT grids need not align perfectly, it must be shown that the impact of misaligned grid
cells is small enough to ignore. For this, the worst-case-scenario of grid misalignment is
considered. For an individual FLUENT cell, the worst case is illustrated in Figure 4.6, where
the cell centroid is aligned with the edge of the MCRT cell. The FLUENT cell is assumed to
be square, with a side length of c.
Since the MCRT cell which contains a given FLUENT cell is determined by the
location of the FLUENT cell centroid, the most overhang a FLUENT cell can have is half its
volume. The most extreme miss-match between the two grid schemes then would be if every
FLUENT cell along every border of a single MCRT cell had this much overhang, and the
FLUENT grid is at its most coarse. The total volume of the overhang along the two b-length
sides can be computed from Equation 4.7.

2 (4.7)
38

Energy History Without Under-Relaxation


Millions 4.80
Net Energy Gain in Receiver [W]

4.75

4.70

4.65

4.60

4.55

4.50
0 5 10 15 20 25 30 35
Iteration Number
Figure 4.5. Net energy gain in the gas flow as a function of model iteration number with
no temperature under-relaxation. This scheme appears to be unstable, or at least not
converging.

Figure 4.6. Illustration of a FLUENT grid cell (in red)


overhanging the boundary of a Monte Carlo grid cell (in
black).
39

Similarly, the total volume of the overhang along the two a-length sides can be computed
from Equation 4.8.

2 (4.8)

Summing these two volumes and dividing by the volume of the MCRT cell yields the total
volume miss-match between the two grid schemes:

% max (4.9)

The largest value of this mismatch occurs in the largest MCRT cell, with the largest
FLUENT cells, where a is 0.5m, b is 0.3m, c is 0.01m, r1 is 1.2m. and r2 is 1.5m, and the
mismatch is 5.33%. In the real grid however, the mismatch is significantly less. On average,
it would be half this, as the FLUENT cells that overhang the edge of an MCRT cell in this
way are equally probable as FLUENT cells that are perfectly aligned with the MCRT cell
edge. Furthermore, many of the FLUENT cells in the outer ring of MCRT cells (i.e. at these
large values of r1 and r2) are made smaller than 0.01m because they are near the wall. It can
also be seen from Equation 4.9 that the worst-case mismatch value also scales linearly with c,
so if further precision is needed it can be had by reducing the maximum grid size of the
FLUENT cells. For regions of high gradients in the receiver, where the grid size is 10-4, this
mismatch value drops to below 0.05%.
40

CHAPTER 5

COLD FLOW & GRID STUDY

5.1 GRID AND GEOMETRY


A schematic of the model receiver geometry is presented in Figure 5.1. Previous work
showed that the gas region between the edge of the aperture and the outer wall of the receiver
experienced minimal heating despite scattered and emitted rays in that region [1]. As a result,
there is little need to model an aperture radius significantly smaller than the receiver radius
itself, thus the receiver geometry has been modified to reduce the radius of the receiver to
that of the window, 1.5 m in this case. Ruthers work also demonstrated the superior
efficiency of flow opposed to the incident radiation [1]. As a result, this models geometry
orients the bulk of the gas flow in the opposed direction. Flow enters the receiver cavity at
the rear wall, travels forward towards the window, and exits down a central outlet tube along
the axis.

Figure 5.1. Schematic of the geometry as it is modeled in FLUENT and the Monte
Carlo code. Drawing is not to scale.
41

In the FLUENT Computational Fluid Dynamics (CFD) portion of the model, the
geometry is two-dimensional-axisymmetric, along the central axis of the cylinder. The
FLUENT geometry must match the geometry in the Monte Carlo exactly, because all source
terms calculated in the MCRT are mapped to their corresponding locations in FLUENT as
described previously in Chapter 4.1. Thus, because the current MCRT model is limited to an
orthogonal axisymmetric geometry, so then must the FLUENT model be likewise limited.
This is the reason for poor aerodynamic features such as the outlet tube entrance and the
corner adjacent to the window, both of which are visible in Figure 5.1. Work to expand the
capabilities of the MCRT solver and relax some of these limitations is ongoing. The grid and
model geometry as seen in FLUENT is shown in Figure 5.2. The axis of the receiver runs
along the lower boundary of the domain, while the window is defined by the left edge. The
black line running through the dense grid cells near the axis represents the wall boundaries of
the outlet tube. The grid sizing is varied throughout the geometry to reduce computational
load while still providing needed resolution in areas of interest. In regions that are both far
from any wall and largely uniform in flow, grid cells are up to 2.5 cm on each side. Cells
near the walls or in areas of high velocity or temperature gradients are on the order of 5mm
on each side.

Figure 5.2. View of the grid as it appears in FLUENT.

The geometry in the Monte Carlo ray tracing radiation model is axisymmetric as well,
although rays are traced in three dimensions. Tracing the rays in three dimensions is required,
because circumferential components of a rays trajectory can change the path length through
the medium as compared to a strictly two-dimensional computation, and thus have an impact
on absorption and scattering calculations. The grid used is nearly identical to that used
42

previously by Ruther [1]. The volume is partitioned uniformly with 10 divisions in the axial
direction and 5 in the radial direction. Figure 5.3 shows the FLUENT grid in green with an
approximate overlay of the Monte Carlo grid shown in black. The MCRT grid is uniform
because of limitations in the original model, which assumes a uniform grid throughout the
code, though a non-uniform grid would be preferable.

Figure 5.3. The FLUENT grid (green) with an approximate overlay of the Monte Carlo
grid (black).

There is an additional axial division at the rear wall of the receiver, to create a set of
five wall elements on that plane with zero thickness. The outer axial wall also has a zero-
thickness radial division, resulting in 10 ring elements at that boundary. These sets of wall
elements are necessary in the MCRT because the ray tracing must have boundaries at these
locations. The back wall elements currently exist only in the MCRT model, because there is
no way to couple them with the FLUENT model at present. The only difference from the
previous model work is the radial dimension of the geometry, which is 1.5 m in this case as
opposed to 2.5m previously. It should be noted that the outlet tube wall does not exist in the
radiation model. Attempts were made to include it, but the modified MCRT code did not
match benchmarks as discussed previously in chapter 2 and thus had to be omitted for this
work. The model therefore behaves as if the outlet tube wall is transparent to all radiation
heat transfer. This is an approximation of a quartz outlet tube, which would be mostly
transparent to incident solar radiation but opaque to longer wavelengths. However, emissions
from the gas within the outlet tube in a real receiver would likely be minimal for two reasons.
First, the gas-particle mixture has a low emissivity in the infrared portion of the spectrum by
design. This is an important feature of the particles selective absorption properties. Second,
43

the gas in much of the outlet tube of a real receiver would have minimal particle content, as
they are expected to oxidize fully by the time the mixture leaves the receiver. Thus the
emissivity of outlet tube gasses is likely to be close to zero, so the omission of the tube wall
from the MCRT is not expected to have a significant impact on overall receiver model
behavior. However, without inclusion of the outlet tube wall in the radiation model, the outlet
tube wall temperatures cannot be accurately predicted. These temperatures will be needed for
the prototype receiver design. Future work will add this to the model now that the problem is
remedied.

5.2 COLD FLOW VERIFICATION


Before coupling the FLUENT model to the MCRT, the FLUENT grid is used to run
an unheated case. For this case, the energy solver in FLUENT is still enabled but the inlet gas
is set to 700 K, all walls are set to the zero heat flux boundary condition, and no source terms
are imposed on the gas volume. All other conditions are the same as the baseline scenario: 5
kg/s mass flow rate inlet, 1 MPa operating pressure, no-slip condition at all walls, and
identical geometry as described previously. The turbulence model used in FLUENT is the
shear-stress transport modified - model with recommended default model parameters.
These parameters are listed in the Appendix. This method was chosen based on FLUENT
user-guide recommendations for internal turbulent flow. The velocity profile at the exit plane
of the outlet tube is compared with a theoretical turbulent velocity profile to validate the cold
flow. Analytical or experimental benchmark cases for the flow in the remainder of the
receiver are not available. The Reynolds number of the flow at the exit plane is roughly
31,000, which is well into the turbulent regime for duct flow. An empirical velocity profile
for turbulent duct flow is given by equation 5.1 in Bejan [26].

2.5 5.5 (5.1)

(5.2)

(5.3)

(5.4)
44

This theoretical profile is plotted in Figure 5.4 against the model results for the cold flow
case at the exit of the outlet tube, for both the - and - turbulence models. Both the
FLUENT models show reasonably good agreement with the theoretical velocity profile, but
the - model does a better job resolving the profile and particularly the boundary layer. The
model under-predicts the turbulent boundary layer thickness slightly compared to theory, but
this is not a focus of the model. Heat transfer across the boundary layer is not a primary
concern of the coupled model, as the dominant modes of heat transfer within the receiver are
radiation, followed by advection within the gas.
6

4
Axial Velocity [m/s]

k-omega (SST)
2 k-epsilon
Theory

0
0.00 0.05 0.10 0.15 0.20 0.25 0.30
Distance from Centerline [m]
Figure 5.4. Plot of the velocity profile on the outlet plane of the receiver for two
FLUENT turbulence models as compared to a theoretical turbulent velocity profile.

Another metric of performance that can be analyzed from the cold flow case is the
pressure drop across the receiver. Figure 5.5 shows contours of static pressure in the receiver,
displayed in Pascals relative to the inlet. The pressure drop from the inlet to the outlet of the
receiver is less than 100 Pa, and the maximum pressure difference is 200 Pa. This extremely
low pressure drop is encouraging, as the receiver geometry is not optimized for aerodynamics
45

Figure 5.5. Cold flow pressure contours in FLUENT.

at all. Additionally, it is much lower than typical pressure drops found in existing tubular
receiver designs. Furthermore, it is well within limits that would be imposed by a gas turbine
system [27]. Figure 5.4 also illustrates that nearly all of the pressure drop occurs in the
entrance region of the outlet tube. Since the current Monte Carlo model requires orthogonal
boundaries and no curved surfaces, the receiver geometry is not optimized for the gas flow.
This includes the outlet tube entrance, which might use a nozzle entrance in a real receiver
and may have a radius which changes in the axial direction.

5.3 GRID STUDY


A grid study is performed using a modified baseline scenario in order to validate grid
independence of that solution and reduce computational load. The solar input conditions for
this case are 5 MW of total power with a Gaussian flux distribution and collimated incident
rays. The gas flow rate is 5 kg/s and all interior receiver walls are modeled with an
emissivity of 1. To reduce processing time, scattering in the gas-particle mixture is turned
off in the radiation model, leaving only emission and absorption. Three grids are tested
under these conditions. Grid 1 has cells as large as 5 cm away from the walls where
gradients are low, and as small as 1 mm near the walls and in the entrance region of the
outlet tube where gradients are highest. Grid 2 is a refinement of grid 1, with every cell
being roughly half as large as in grid 1. Grid 3 is a refinement of grid 2, with each cell again
46

being half as large as the previous grid 2. Table 5.1 shows a summary of the differences
between each of the three grids.

Table 5.1. Summary of Grid Attributes


Grid 1 Grid 2 Grid 3
Number of Cells 150,000 300,000 600,000
Smallest Cell [m] 0.001 0.0005 0.00025
Largest Cell [m] 0.05 0.025 0.001

Figure 5.6 shows a plot of temperatures on the outer cylinder wall of the receiver for
each of the three grids. Each case was run for 105 iterations with a ramping emissions
scheme as described previously in Section 4.1.

GridStudy CylinderWallTemperatures
1100

1050 grid1
grid2
1000 grid3

950
WallTemperature[K]

900

850

800

750

700

650

600
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0
AxialDistancefromWindow[m]
Figure 5.6. Wall temperatures in FLUENT for each of the three grids studied.
47

Temperatures in the first 0.5 m are variable and unreliable due to the unstable nature
of the flow field in the corner adjacent to the window. In spite of this however, even the most
extreme temperatures are only 1050 K which is manageable for a variety of potential wall
materials. Grids 2 & 3 show a quicker return to expected wall temperatures between 0.5 and
1 m than grid 1 does. The reason for this can be seen in a comparison of the velocity fields
near the corner for each grid. Figure 5.7 shows velocity vectors near the corner of the
window and outer cylinder wall for grid 1. Vectors are colored and scaled to velocity
magnitude. The window is on the left side, the outer cylinder wall is along the top, and the
white vertical line near the right side denotes a distance of 0.5 meters from the window face.
A large recirculation cell is visible in the corner, extending roughly 0.3 meters from the
window. A second smaller recirculation cell can be seen near the wall extending past the 0.5
meter mark. It is this smaller recirculation cell which causes the irregular wall temperatures
seen previously in Figure 5.7 for grid 1, between 0.5 and 1.0 meters.

Figure 5.7. Velocity vectors in the window corner region for Grid 1. Vectors are colored
and scaled by magnitude. Color-bar scale represents velocity in m/s, and length scale of
the geometry is indicated at the top of the figure.

Figure 5.8 and Figure 5.9 show velocity vectors in the same corner region for grids 2
& 3, respectively. A recirculation cell still exists in the corner, but it is smaller than the one
seen in grid 1, and the flow along the wall remains attached closer to the window than under
48

Figure 5.8. Velocity vectors in the window corner region for Grid 2. Vectors are colored
and scaled by magnitude. Color-bar scale represents velocity in m/s, and length scale of
the geometry is indicated at the bottom of the figure.

Figure 5.9. Velocity vectors in the window corner region for Grid 3. Vectors are colored
and scaled by magnitude. Color-bar scale represents velocity in m/s, and length scale of
the geometry is indicated at the top of the figure.
49

the coarser grid scheme. This leads to stable wall temperatures closer to the window for grids
2 and 3, as seen previously in Figure 5.6. There are still some differences between the flow
fields seen in grids 2 and 3, such as the size of the largest recirculation cell and the
magnitude of the velocity in that region. But these differences are not significant enough to
justify the increased computational load needed for the extremely fine grid 3.
Figure 5.10 and Figure 5.11 provide some explanation for the unusual wall
temperatures found in the first 50 centimeters of the receiver, and illustrated previously in
Figure 5.6. Figure 5.10 shows temperatures in this wall region for grid 2, while Figure 5.11 it
shows velocity vectors in the same region. The position axis of Figure 5.10 is scaled to align
with the image in Figure 5.11. The velocity vectors are scaled by magnitude, but colored by
flow temperature. The source term imposed on the wall from the Monte Carlo in this region
is negative, meaning the wall is a net emitter of radiation and thus is heated by the flow.
1200
Wall Temperature [K]

1100

1000

900

800

700

600
0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40 0.45 0.50
Axial Distance from Window [m]
Figure 5.10. Cylinder wall temperatures in the first half-meter of the receiver. The x-
axis is scaled to align with the velocity vectors in Figure 5.11.

From these two plots, it can be seen that the wall temperature drops in regions where
the flow separates from the wall, specifically in three locations: immediately adjacent to the
window, 15 cm away from the window, and 30 cm away from the window. In regions where
the flow is relatively faster and attached to the wall, temperatures are higher because
convection is more effective due to the higher flow velocity. As the flow field in this corner
region changes due to grid resolution or possible transients, so does the wall temperature
50

Figure 5.11. Velocity vectors in the window corner region for Grid 2. Vectors are scaled
by magnitude of velocity, and colored by temperature in Kelvin.

profile.
At this point in the analysis of wall temperature and heat flux, it is appropriate to
justify the model assumption that conduction heat losses through the wall to the ambient are
small enough to be neglected. Equation 5.5 shows the energy balance on the outer cylinder
wall, where G is the irradiance on the wall, the second term represents emitted radiation from
the wall, the third term represents convection from the gas adjacent to the wall, and
represents the radial conduction losses through the wall to the ambient. As discussed
previously in Chapter 4.2 and illustrated in Figure 4.2, the magnitude of the heat flux on the
wall due to radiation is limited to 5 kW/m2. This is a net radiation flux, so it is the difference
between the first two terms of Equation 5.5. The FLUENT model then assumes that the
term is zero, leaving the convection term as the only unknown.

0 (5.5)

For conduction losses to be negligible, they should be less than 5% of the net
radiation flux, or less than 200 W/m2. This flux value can be converted to a total heat loss
term by multiplying by the outer receiver wall surface area, which yields 9450 W or roughly
10 kW. This is 0.2% of the total power absorbed by the receiver, showing that the conductive
losses can be made to have a small effect on receiver efficiency. The conduction heat loss
51

through a cylinder wall is expressed by Equation 5.6 [28], where the previously computed
value of 10 kW is represented by , L is the length of the receiver, k is the thermal
conductivity of the insulation around the receiver body, Tw is the inner wall temperature of
the receiver, To is the exterior wall temperature of the insulation, r2 is the outer radius of the
insulation, and r1 is the radius of the receiver wall.
2 (5.6)
ln
A conservative estimate of average receiver wall temperature is 1000 K, and the outer
temperature of the insulation can be taken to be approximately equal to an ambient
temperature of 300 K for a worst-case estimate of the heat loss. The radius of the receiver is
fixed at 1.5 m and the length is fixed at 5 m. A conservative estimate of the average thermal
conductivity of the high-temperature insulating material able to withstand temperatures of
1000K is 0.15 W/(m K) [29]. Solving equation 5.6 above for r2 gives an outer insulation
radius of 2.1m, or an insulation thickness of 60 cm. This is a reasonable worst-case scenario
for needed insulation thickness, and could be further reduced by using a second insulation
material with reduced thermal conductivity at the outer radius ranges where temperatures are
reduced. For example, using the high-temperature insulation for 15 cm radially outward from
the wall would reduce the temperature to 800 K, where a insulation with an average thermal
conductivity of 0.07 W/(m K) could be used [28]. This second layer would need to be 20 cm
thick to reach the ambient temperature, yielding a total insulation thickness of 35 cm.
Returning to the grid study, another metric used to compare grid schemes is the net
radiation heat transfer along the outer cylinder wall. Figure 5.12 shows the wall flux values
applied in FLUENT for each of the three grids. These values are generated in the MCRT
code based on the result of the radiation heat transfer calculations between the walls and the
volume. There is good agreement between grids 2 and 3 for the wall flux profile, while grid 1
fluxes stand apart, particularly at each end. Wall flux values are negative near the front of the
receiver, increasing towards the rear. With a collimated solar at the window, the wall
elements near the front of the receiver do not experience any direct solar irradiation. They are
predominantly irradiated by re-emissions from the gas itself near the wall. The amount of
light illuminating the gas near these early wall elements is itself reduced by the Gaussian
input flux distribution, which imparts only 100 kW/m2 on the outer most ring element of the
52

Grid Study - Wall Fluxes

5000
grid1
4000 grid2
grid3
3000
NetWallFlux[W/m2]

2000

1000

1000

2000

3000
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0
AxialDistancefromWindow[m]
Figure 5.12. Wall fluxes imposed by the Monte Carlo on the FLUENT solver for each of
the three grids studied. Positive values of this net wall flux means the wall is absorbing
more radiation than it is emitting.

aperture. Wall elements near the middle of the receiver experience radiation emissions from
gas regions in a wide band of axial locations, while wall elements near the front of the cavity
receive no gas emissions from axial positions to their left. This reduces the magnitude of gas
emissions that reach the wall elements near the window. An additional source of radiative
load that acts on the wall elements near the rear of the cavity more than the window end is
the back wall of the receiver. This wall exists in the Monte Carlo radiation model as a
black surface and much of what it emits is absorbed by the cylinder wall. That wall reaches
temperatures of up to 1100K in the Ruther model.
The final comparison between the grids is the average outlet temperature of gas
leaving the receiver. All three grid schemes produced very similar outlet temperatures of
1442 K, 1443 K, and 1445 K for grids 1, 2, and 3, respectively. As the outlet temperature of
53

the gas is the most important benchmark of the receiver model, it is significant that all three
grids produced such close results in this regard. Grid 2 was chosen for model study moving
forward because the differences between grids 2 and 3 were minimal compared to the
differences between grid 1 and the rest. Grid 2 was chosen over grid 3 because the small
improvements to accuracy did not outweigh the increased computational load of the finest
grid.
54

CHAPTER 6

RESULTS AND DISCUSSION

Once the model was developed and verified, three different parameters were selected
for further study. These are the influence of Gaussian flux distribution as compared to a
uniform solar flux input, the effects of variation in mass-flow rate through the receiver with a
constant solar input, and the effects of varying mass-flow rate proportionally with a varying
flux input. Ten cases in all were conducted to examine these effects, and their differences are
summarized in Table 6.1, along with macro-scale results. For the flux distribution variation,
two cases were run at the baseline conditions of 5 kg/s mass-flow rate and 5 MW of
collimated solar input. For the mass-flow rate study, five cases were run with a 5 MW solar
input and mass-flow rates from 4.0-6.0 kg/s. For the variable input study, five cases were run
with input power varied from 2.0-6.0 MW and mass-flow rate varied proportionally from
2.0-6.0 kg/s.

Table 6.1. Summary of Model Runs


Mass Net
Input Window Outlet
Gaussian Flow Gain Thermal
Power Losses Temperature
Flux? Rate (MCRT) Efficiency
[MW] [kW] [K]
[kg/s] [MW]
5 no 5.0 812 4.188 1379 83.8%
5 yes 4.0 683 4.317 1547 86.3%
5 yes 4.5 587 4.413 1479 88.3%
5 yes 5.0 538 4.462 1420 89.2%
5 yes 5.5 509 4.491 1370 89.8%
5 yes 6.0 477 4.522 1325 90.4%
6 yes 6.0 597 5.403 1424 90.1%
4 yes 4.0 485 3.515 1409 87.9%
3 yes 3.0 451 2.549 1389 85.0%
2 yes 2.0 393 1.606 1342 80.3%

For all cases, several parameters of the model were held constant. These include a
particle size of 0.2 m and particle mass loading of 0.30 kg/m3, along with the flow
55

orientation being kept predominantly opposed to the incident flux. These parameters were
studied in more detail with the original version of the model by Ruther [1], and these values
were chosen based on those results. The boundary conditions are also held constant for all
cases, with the exception of the solar input for certain studies. Constant boundary conditions
include the no-slip condition at all walls, a uniform inlet velocity profile, 700 K constant
temperature inlet gas, and a constant temperature of 1000 K at the window wall. The window
boundary condition was chosen to be a constant temperature because a quartz window in a
real receiver will have a temperature limit near this. By forcing the window temperature to be
1000 K, the computed heat flux at that boundary in FLUENT gives an idea of the convective
heating load on the window. In most cases, this flux was on the order of 50-100 kW. In a real
receiver, the convective load would be far outweighed by the radiative load on the window,
which would largely come in the form of longer wavelength emissions from the gas and
walls. Longer wavelengths of course are likely to be absorbed by a quartz window. Since the
current Monte Carlo model does not include an actual quartz window at the aperture, there is
no way to include radiative loads on the window as of yet, though their magnitude can be
estimated by the window losses shown in Table 6.1. Thermal loads on the window and
potential strategies for reducing them are being investigated by other members of the
research group.

6.1 GAUSSIAN FLUX DISTRIBUTION


As discussed previously in Chapter 3.1, the Monte Carlo model has been modified to
allow for a Gaussian flux distribution on the aperture. To assess the impact of this change on
the receiver, the baseline coupled case with 5 kg/s mass-flow rate and 5 MW of solar input
was run with a uniform flux distribution and again with the new Gaussian flux distribution.
Figure 6.1 shows the energy balance over the model runs for both of these cases, with the
uniform flux case in blue and the Gaussian flux case in red. The energy balance is computed
by comparing the total energy gained by the flow in FLUENT to the net energy gained in the
Monte Carlo model. The energy gain in the Monte Carlo is simply the input power (5 MW in
this case) less all reported window losses, which are all rays which are traced to exit the
aperture plane. For both these cases (and indeed nearly all coupled cases presented in this
56

Energy Balance - Gaussian Flux Distribution


102%

101%
Energy Balance (FLUENT/MCRT)

UniformFlux
100%
GaussianFlux
99%

98%

97%

96%

95%

94%

93%

92%
0 10 20 30 40 50 60 70 80 90 100 110
Iteration Number
Figure 6.1. Effect of solar flux distribution on energy balance.

thesis) the energy balance is several percent less than 100%. Potential sources of this
imbalance are discussed further later in this chapter.
An important feature illustrated in Figure 6.1 is the fact that the uniform flux case
produced an energy balance of 97%, while the Gaussian flux case is less than 96%. One
potential reason for this difference is evident when looking at the temperature fields for each
case, which are shown in Figure 6.2 and Figure 6.3. In the uniform flux case, the peak gas
temperature in the receiver cavity is around 1750 K, seen near the entrance to the outlet tube
in the lower-left corner of Figure 6.2. This value is well below the model-enforced maximum
temperature of 2200 K, and thus FLUENT did not artificially reduce the gas temperature in
any cells for the uniform flux case. By contrast, the maximum gas temperature found in the
Gaussian flux distribution case is 2200 K, visible along the axis near the window. Since the
gas temperature in these cells is limited in FLUENT, some of the energy that the Monte
Carlo solver determined ought to be gained in this region is simply ignored by the model. In
57

Figure 6.2. Temperature field inside receiver with 5 MW of uniform incident solar flux
and 5 kg/s mass-flow rate. Color-map values are in units of Kelvin.

Figure 6.3. Temperature field inside receiver with 5 MW of Gaussian distributed


incident solar flux and 5 kg/s mass-flow rate. Color-map values are in units of Kelvin.

other words, if all of the energy sent from the MCRT to FLUENT was accounted for, these
gas cells would have temperatures much higher than 2200 K, and thus the average outlet
temperature of the flow would be higher. This results in an inherent imbalance in the energy
gained by the gas as reported by the two models, and explains the difference in energy
balance found in the uniform and non-uniform input flux cases. In a real receiver, gas at this
temperature would oxidize the small carbon particles extremely quickly, thus making the gas
transparent to the incident radiation and unable to continue gaining energy in this way. But
58

because the current model does not have a way to account for oxidation of the small carbon
particles, this controlling effect is not captured.
This gas temperature is extremely high for several reasons. First, the receiver axis is
the peak of the Gaussian flux distribution, and thus receives nearly 3 MW/m2 on the aperture
at this radial position. Second, there is a stagnation point in on the window boundary at the
axis of symmetry, which limits the ability of the heat here to be carried away by advection in
the flow. This stagnation point is a result of the orthogonal and axisymmetric geometry
restrictions of the current model. The stagnation point is also responsible for the relatively
cooler gas in this same region seen in the uniform flux case. In that case, the cooler gas is in
part a result of the constant temperature boundary condition of 1000 K imposed at the
window plane in FLUENT.
Another significant difference between the two flux distribution cases is the
temperature distribution within the receiver. The uniform flux case has a larger region of
relatively hot gas (1600-1750 K) in the front half-meter of the cavity, while the non-uniform
flux case has gas at this temperature only near the axis. This is again a result of lower flux
levels at the outer radius in the second case. This also produces a slight radial temperature
gradient in the non-uniform flux case that persists nearly all the way to the back of the cavity,
while the uniform flux case has minimal radial temperature change in the bulk of the gas
volume. The hot gas in the front of the uniform flux case produces higher radiation emission
levels in that region than its analog in the Gaussian distribution case. This in turn results in
greater window losses computed by the Monte Carlo solver, as much of the emissions from
hot gas near the front of the cavity simply exit the aperture. Consequently, the uniform flux
case has a lower thermal efficiency (83.9%) than the non-uniform case (89.2%), as well as a
lower outlet temperature (1379 K vs. 1420 K).
An additional source of losses for the uniform flux case is apparent when examining
the outer cylinder wall temperatures for these two model runs. Figure 6.4 shows the wall
temperature profiles for these two variations. Wall temperatures are clearly higher in the
uniform flux case, particularly in the front portion of the receiver. This result is expected, as
incident flux levels are much higher near the wall in the uniform case and in turn more
incident radiation is available to be scattered by the gas-particle mixture and strike the wall.
Since the wall is modeled as a black body, it must re-emit nearly all of the energy absorbed.
59

Wall Temperature Profiles


1300

1200
Cylinder Wall Temperature [K]

UniformFlux

1100 GaussianFlux

1000

900

800

700
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0
Axial Distance from Window [m]
Figure 6.4. Cylinder wall temperature profiles for the uniform (blue) and Gaussian
(red) flux distribution cases.

And since the wall temperature is roughly between 900 and 1200 K, the bulk of its emissions
are at longer wavelengths where the gas-particle mixture does not absorb very well. This
means that most of the emissions from the outer cylinder wall will exit the aperture and be
realized as losses, or be re-absorbed by the wall in another location. Only a limited amount of
the walls energy is removed via convection by the flow, as was discussed previously in
Chapter 4.2.
All of these factors together make the Gaussian flux case preferable for receiver
performance from a design standpoint, in addition to this flux distribution being more
realistic than a uniform case. Although the model is still limited in the variety of geometries
it can solve, some of the trends illustrated by these two cases are likely applicable to non-
orthogonal receiver designs. Keeping incident solar flux levels low near the cylinder wall,
60

even with a collimated radiation input, likely improves thermal efficiency of the receiver and
reduces radiative load on the wall itself. Likewise, keeping the hottest gas region limited to a
smaller volume of fast flowing gas near the entrance to the outlet tube helps to reduce
radiation losses and thus improve receiver thermal efficiency as well.

6.2 FLOW RATE VARIATION


To study the effects of mass-flow rate through the receiver, two cases each were run
above and below the baseline mass-flow rate of 5.0 kg/s, for a total of five different mass-
flow rates with a constant 5 MW of solar input as summarized in Table 6.1. Results of outlet
gas temperature and receiver thermal efficiency for each of the five cases are plotted in
Figure 6.5.

Gas Outlet Temperature vs. Mass Flow Rate


1600 91%

90%
1550 89%
Outlet Temperature [K]

88%

Receiver Efficiency
1500
87%

86%
1450
85%

84%
1400 Temperature
Efficiency 83%

1350 82%

81%

1300 80%
3.0 3.5 4.0 4.5 5.0 5.5 6.0 6.5 7.0

Gas Mass Flow Rate [kg/s]

Figure 6.5. Effect of mass-flow rate on gas outlet temperature and receiver thermal
efficiency, with 5 MW collimated Gaussian input.
61

As might be expected, the outlet gas temperature is inversely proportional with the
mass-flow rate through the receiver, at least over this range of mass-flow rates near the
baseline scenario. Outlet temperatures range from 1325 K to 1550 K over a fairly narrow
window of mass-flow rates. This provides a potential design criterion for a real receiver, as
flow rates can be adjusted to reach a target outlet temperature for downstream needs such as
thermal storage or a turbine. Figure 6.6 puts the result from Figure 6.5 into the context of
power plant efficiency and currently existing receivers. Figure 6.6 plots receiver, engine, and
combined efficiency against operating temperatures [2]. For the receivers, operating
temperature refers to outlet temperature, while for the heat engine curve it refers to turbine
inlet temperature. The blue line shows model predicted efficiency of the SDSU small particle
receiver as a function of outlet temperature, at an input power of 5 MW. The red line shows
similar data for existing tubular receivers, which operate at lower temperatures than the small
particle receiver. The green curve shows heat engine efficiency for existing power plants, and
the purple curve shows the combined efficiency of existing receivers & existing heat engines.
Currently we do not have a calculation for the small particle receiver coupled to a suitable
gas turbine to show a combined line for that combination, but that is the subject for future
work. We can conclude that the small particle receiver results show significant improvement
compared to current receiver technology. The ability to maintain receiver efficiency above
85% at higher temperatures allows for higher engine efficiency and thus greater overall
efficiency of the solar thermal power plant.
Figure 6.5 also shows that efficiency improves as mass flow rate goes up over this
range, although the gains are reduced as efficiency approaches 90% or so. This trend is likely
the result of a similar mechanism to that seen in Chapter 6.1. Specifically, lower mass flow
rates result in higher gas temperatures in the receiver, which in turn result on greater
radiation losses out the aperture. This can be seen not only in the outlet temperatures as
shown previously in Figure 6.5, but also in the temperature field. Figure 6.7 and Figure 6.8
show the temperature field inside the receiver at the two extreme cases of 4.0 and 6.0 kg/s
mass-flow rates, respectively. While both cases have regions of model-limited temperatures
of 2200 K near the axis and the window as previously explained in Chapter 6.1, this region is
larger in the lower mass-flow case. Additionally, the gas in the front portion of the receiver
outside of this temperature-limited zone is hotter in the lower mass-flow case than in the
62

Impact of Receiver Efficiency on Power Plant


100%

90%

80%
SPR@5MW
70%
ExistingReceivers
60% EngineEfficiency
Efficiency

Total(existing)
50%

40%

30%

20%

10%

0%
400 600 800 1000 1200 1400 1600 1800
OperatingTemperature[K]
Figure 6.6. Receiver, engine, and total efficiencies as a function of operating
temperature.

Figure 6.7. Temperature field in the receiver with a mass-flow rate of 4.0 kg/s and input
power of 5 MW. Color-map numbers are in units of Kelvin.
63

Figure 6.8. Temperature field in the receiver with a mass-flow rate of 6.0 kg/s and input
power of 5 MW. Color-map numbers are in units of Kelvin.

higher mass-flow case. The mass-weighted average temperature in the hottest Monte Carlo
cell (within 0.5 m of the window and 0.3 m of the axis) is 1661 K for the 4.0 kg/s case, as
compared to 1326 K for the 6.0 kg/s case. Thus, this cell emits more radiation in the lower
mass-loading case, much of which exits the receiver aperture and is tallied as a radiative loss.
This trend persists moving radially outward, resulting in overall greater radiative losses for
the lower mass-loading case, thus reducing receiver efficiency.
This issue is also apparent when we examine the energy balance for each of the five
cases. Figure 6.9 shows the energy balance for each case as a function of iteration number.
The colder cases show a better energy balance. The improved energy balance could be a
result of the same mechanism discussed in the previous chapter, that being the relative size of
model-limited temperature regions in the gas. The higher mass-flow rate cases have smaller
volumes of gas at the model limit of 2200 K, thus less of the energy input from the Monte
Carlo model are neglected in the FLUENT model for these cases. Furthermore, it is likely
that the true temperatures in this 2200 K region are higher in the low mass-flow cases. This
would result in a greater magnitude of energy neglected by the FLUENT solver. The lower
mass-flow rate cases also show less stability in the energy balance during iteration. This is
possibly due to temperature swings but an explanation has not yet been determined.
Another difference between mass-flow rate cases, visible in Figure 6.7 and Figure 6.8
is the radial temperature gradient in the outlet tube. The lower mass-flow rate case has a
significant temperature gradient which persists all the way to the exit plane. In contrast, the
64

Energy Balance - Mass Flow Rate Variation


102%

101%
6.0kg/s
100% 5.5kg/s
Energy Balance (Fluent/MCRT)

5.0kg/s
99% 4.5kg/s
4.0kg/s
98%

97%

96%

95%

94%

93%

92%
0 10 20 30 40 50 60 70 80 90 100 110
Iteration Number
Figure 6.9. Energy balance as a function of iteration number, for five different mass-
flow rate cases, 5 MW input power.

higher mass-flow rate case experiences more turbulent mixing and thus has a nearly uniform
gas temperature by the time the flow reaches the outlet plane. This is not an issue of serious
concern, as downstream piping could be used to induce more mixing if needed.
As a result of lower gas temperatures in the higher mass-flow cases, wall
temperatures were also reduced as mass-flow rate increased. This trend can be seen in Figure
6.10, which plots the outer cylinder wall temperature profile for each of the five cases.
Again, as with the wall temperature differences discussed previously in Chapter 6.1, reduced
wall temperatures lead to reduced window losses and greater receiver efficiencies, since
much of the emissions from the black walls exit the aperture. This is especially true for wall
elements near the front of the receiver.
65

Wall Temperatures - Mass Flow Rate Variation


1300

6.0kg/s
1200 5.5kg/s
5.0kg/s
4.5kg/s
Wall Temperature [K]

1100 4.0kg/s

1000

900

800

700
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0
Axial Distance from Window
Figure 6.10. Axial cylinder wall temperature profiles for five different mass-flow
rates, 5 MW input power.

6.3 INPUT POWER VARIATION


To study the effect of varying input power to the receiver, as well as to explore a
potential control mechanism for the receiver, five cases were run with differing mass-flow
rates and input power levels. As summarized at the beginning of this chapter and in Table
6.1, the ratio of mass-flow rate to input power is held constant for these cases at the same
level as the baseline scenario. On the low end, this gives a mass-flow rate of 2.0 kg/s and
input power of 2 MW, while on the high end these values are 6 kg/s and 6 MW respectively.
This is done in an attempt to model using mass-flow rate variation to regulate receiver
behavior under varying input power levels. Because heliostat fields produce variable power
levels over the course of a day and a year due to changing solar flux levels, it may be
desirable to change receiver operating conditions in response. One potential motivation for
doing so would be to maintain a relatively constant outlet temperature to feed downstream
66

processes such as a turbine or thermal storage system. Results for outlet gas temperature and
receiver thermal efficiency are presented in Figure 6.11. While temperatures were higher at
higher input levels, the range of outlet temperatures seen here is relatively narrow, ranging
from 1342 K on the low end to 1424 on the high end. Furthermore, power levels between 4
and 6 MW produced outlet temperatures that differ by only 25 degrees Kelvin. This is a
remarkably consistent output considering that input power levels increased by 50% in this
range. Finer tuning of the mass-flow rate variation may provide the ability to maintain nearly
constant outlet temperatures over a broad range of input power levels.

Outlet Temperature & Thermal Efficiency


1450 92%

1430 90%
Outllet Temperature [K]

1410 88%

Thermal Efficiency
1390 86%

1370 84%
Outlet Temperature
Thermal Efficiency
1350 82%

1330 80%
1 2 3 4 5 6 7
Mass-Flow Rate [kg/s] & Inlet Power [MW]
Figure 6.11. Gas outlet temperatures and receiver thermal efficiencies for five different
inlet power levels.

This control mechanism may come at some cost to efficiency however, as Figure 6.11
also illustrates that receiver efficiency went down as power levels and mass-flow rates were
reduced. At the low end of input power and mass-flow rate, the receiver efficiency was
barely above 80%, while at the upper end receiver efficiency was over 90%. The reduced
67

receiver efficiency at lower mass-flow rates is attributable to the size of the receiver relative
to the input power level. Though the input power is reduced, the aperture size remains
constant and thus the potential radiative losses remain relatively constant. The temperature
fields for these cases shows minimal differences. Figure 6.12 and Figure 6.13 show the
temperature field in the receiver for the 2 kg/s and 6 kg/s mass-flow rate cases, respectively.
While outlet temperatures are similar between the two, the higher mass-flow rate case has a
small gas region at the model-limited temperature of 2200 K, while the peak temperature in
the lower mass-flow rate case is barely over 1600 K. Despite this, the lower mass-flow case
has a greater mass-weighted average temperature in the hottest Monte Carlo cell by about 50
degrees Kelvin. This trend continues to some extent as we move radially outward, thus
increasing emissions from the gas regions adjacent to the aperture and increasing window
losses accordingly.

Figure 6.12. Temperature field inside the receiver with a mass-flow rate of 2 kg/s and
input power of 2 MW. Color-bar units are Kelvin.

Differences in wall temperatures also provide little explanation for the reduced
efficiency at lower power levels. Figure 6.14 shows axial wall temperature profiles for the
five cases of variable inlet power. Though wall temperatures are all within a relatively
narrow band, the lower power level and mass-flow rate cases actually show a trend towards
lower wall temperatures, particularly in the front half of the receiver. As was the case in the
previous chapter, higher wall temperatures, especially near the aperture, produce greater
radiative losses and thus drive down receiver efficiency. Despite the slightly larger radiative
68

Figure 6.13. Temperature field inside the receiver with a mass-flow rate of 6 kg/s and
input power of 6 MW. Color-bar units are Kelvin.

Wall Temperature Profiles - Input Power Variation


1100

1050

1000
Wall Temperature [K]

950

900

850

2kg/s 2MW
800 3kg/s 3MW
4kg/s 4MW
5kg/s 5MW
750 6kg/s 6MW

700
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0
Axial Distance from Window [m]
Figure 6.14. Axial cylinder wall temperature profiles for five different inlet power
levels and mass-flow rates.
69

losses from the wall in the high power and high mass-flow rate cases, total radiative losses
are still greater in the low power cases. This is a feature that may be important to keep in
mind when designing a real receiver. Although the model has shown wall temperatures that
are relatively low from a material strength standpoint, keeping wall temperatures down and
reducing emissions from the walls can help reduce losses and increase efficiency of the
receiver. Additionally, the receiver size and in particular aperture size must be carefully
designed for a relatively narrow range of power inputs. Decreasing power by a factor of
three, as is the case when going from 6 MW to 2 MW, forces the receiver to operate far from
its design point and thus at reduced efficiency.
70

CHAPTER 7

CONCLUSIONS AND FUTURE WORK

7.1 CONCLUSIONS
The coupling of the Monte Carlo radiation solver to the FLUENT flow solver brings
the model closer to a realizable geometry. On balance, the move to more realistic flow
geometry including the outlet tube did not change receiver efficiency significantly from the
results of the original model by Ruther, but the current model provides more confidence in
the results and will greatly help with actual receiver design such as sizing the outlet tube,
determining wall and window temperatures, and finding the pressure drop through the
receiver. Efficiencies remained between 80-90%. Outlet gas temperatures remained in the
range of 1300 K to 1550 K depending on input conditions. While attempts to include the
outlet tube wall in the Monte Carlo ray-trace solver were not successful until after the
completion of this research, the challenge has recently been overcome and future work will
include it, as discussed in Chapter 7.2.
The introduction of a Gaussian flux distribution on the aperture not only matched
reality better than the previous uniform flux case, but also improved receiver efficiency by
about 5% and increased outlet gas temperature by 40 degrees Kelvin. This improvement
highlights the importance of keeping the bulk of incident radiation focused in the gas volume
and away from the walls. A real receiver design should take this into account when
determining receiver wall geometry, perhaps by re-introducing a slight offset between the
aperture and the outer wall, as well as by angling the outer wall to align parallel with the
most extreme angle of incident radiation. Wall temperatures and radiation losses may also be
kept down by utilizing a more aerodynamic geometry and eliminating sharp corners which
cause recirculation. The recirculation cells which developed in the corner in all cases
produced slightly elevated gas temperatures, which contributed to window losses and made
wall temperatures more difficult to predict.
The model demonstrated that mass-flow rate may be used effectively to adjust
receiver outlet temperature. However, this control mechanism also impacts receiver
71

efficiency significantly. A 50% increase in mass-flow rate from 4 kg/s to 6kg/s reduced
outlet temperatures by more than 200 degrees Kelvin, but increased receiver efficiency by
4%. The highest mass-flow rate modeled at 6 kg/s also produced the greatest receiver
efficiency, which suggests that further investigation of higher mass-flow rates is warranted.
For the first time, a graph of receiver efficiency vs. output temperature was produced that
allows this receiver to be compared to others that might be used to run a gas turbine, as well
as to determine the optimum coupling temperature as shown in Figure 6.6.
Variations in input power along with mass-flow rate showed how sensitive the
receiver design can be to input flux conditions. Reducing the input power level (and therefore
flux concentration) well below the design point of 5 MW significantly reduced receiver
efficiency. This is likely due to radiation losses not being similarly reduced in conjunction
with the reduction in incident flux. It is therefore important to size a real receiver aperture as
small as possible for the intended heliostat field. Additionally, the trend suggests that this
receiver may benefit from an even greater peak flux than that investigated here, which was 3
MW/m2.

7.2 FUTURE WORK


There are many avenues open to improving the current receiver model. The most
immediate improvement, which is currently underway, is the addition of the outlet tube wall
to the Monte Carlo ray-trace model. Preliminary results suggest that the impact of this
addition will not significantly affect receiver outlet temperatures or efficiencies. However, it
is important to include the feature because an estimate of outlet tube wall temperatures is
needed for the design of a real receiver. Emission from this solid surface may put a local
heating load on the window as well that is currently not accounted for.
A second useful change to the model would be a shift to a non-uniform grid in the
Monte Carlo radiation solver. A non-uniform grid in both the axial and radial directions
could be used to provide greater resolution near the front and center of the receiver
respectively, where temperature gradients are higher than in the rear and near the outer edge.
This could be accomplished without increasing the total number of MCRT grid cells, and
thus without a major impact on computation time. Along with adjustments to the MCRT grid,
the FLUENT grid could be refined based on the results presented in this thesis. In particular,
72

finer grid resolution near the walls is needed to fully resolve some of the thermal boundary
layers in the gas. If the refinement is limited to just the first few cells near the wall, it would
not impact computational load significantly. Conversely, it may be possible to further
coarsen the grid in the volume away from the walls to offset the increased load of refinement
near the walls.
The addition to the code likely to impact results of the model most significantly is a
coupled particle oxidation model. As the gas-particle mixture heats up, eventually the small
carbon particles will oxidize into CO2 and CO. As they oxidize, their size will be reduced
until finally the particles disappear altogether. This has a significant effect on the absorption
and emission characteristics of the gas-particle mixture in the receiver. Gas regions with fully
oxidized carbon would have no particles and thus be virtually transparent to incident
radiation. This is expected to eliminate the small overheated gas region found near the center
of the window in many cases, and may also act somewhat as a control mechanism in the
receiver. It may also reduce absorption in the outlet tube gas, which may reduce receiver
efficiency or warrant a modification to the outlet tube geometry. Modeling particle oxidation
in conjunction with the radiation energy exchange and flow solution is a significant
undertaking which may require three-way coupling with a new code. It is also likely to
increase computation time still further.
Another important feature that is still needed in the model is the ability to model non-
orthogonal boundaries in the Monte Carlo. This would allow for the inclusion of a curved
window, as well as an angled or curved outer wall. A real receiver will require a curved
window to contain pressure [21, 30], and these two features will likely have an impact on the
flow field in the receiver. This will require significant changes to the Monte Carlo code,
possibly including a shift to fully three-dimensional modeling, and likely will increase
processing time as a result. It may be necessary to re-structure the Monte Carlo code to allow
for parallel processing to reduce computation time. Should the code transition to a fully
three-dimensional model, natural convection and the effect of swirl velocity should also be
investigated.
Currently, only the radiation flux values applied to the outer cylinder wall are
interpolated, because sensitivity of the solution demanded it and the interpolation scheme
was relatively simple being only one-dimensional. Expanding the source term interpolation
73

scheme to include two-dimensional interpolation of volumetric source terms in the gas would
improve accuracy of the temperature field results. Though the impact may not be all that
significant in the aggregate, local changes might provide important insight. This change
would be fairly straightforward to implement in the current geometry, but would become
significantly more complex with a non-uniform grid, non-orthogonal boundaries, or a fully
three-dimensional Monte Carlo solver.
Aside from expansions to the model, the existing code could be used to examine a
wider range of input parameters including non-collimated incident radiation and a wider
range of input powers and mass-flow rates. It would also be beneficial to scale the model
down to match the size and approximate geometry of a lab-scale receiver currently being
built and tested at San Diego State University [30]. This would serve as an important
validation of the model, and help to quantify the importance of some current approximations.
A wide range of input parameters could then be used in conjunction with an overall cycle
model previously written by Kyle Kitzmiller [27] to characterize system performance with
the receiver coupled to a gas turbine. This would allow for a more robust version of the graph
in Figure 6.6 which showed receiver and combined efficiency as a function of operating
temperature.
74

REFERENCES

[1] Ruther, Steven James. Radiation Heat Transfer Simulation of a Small Particle Solar
Receiver Using the Monte Carlo Method. Masters Thesis, San Diego State
University, 2010.
[2] Stine, William B., and Michael Geyer. Central Receiver Systems. Power from the
Sun. Last modified October 2012. http://www.powerfromthesun.net/Book/chapter10/
chapter10.html.
[3] NREL. Gemasolar Thermosolar Plant. Concentrating Solar Power Projects. Last
modified October 24, 2011. http://www.nrel.gov/csp/solarpaces/project_detail.cfm/
projectID=40.
[4] Torresol Energy. Gemasolar. Gemasolar.jpg. Last modified June 21, 2011.
http://en.wikipedia.org/wiki/File:Gemasolar.jpg.
[5] Stine, William B., and Michael Geyer. Power Cycles for Electricity Generation.
Power from the Sun. Last modified October 2004. http://www.powerfromthesun.net/
Book/chapter12/chapter12.html#12.4 Brayton Cycle Engines.
[6] Siemens. Trail Blazing Power Plant Technology. Fossil Power Generation Division.
Last modified May 19, 2011. http://www.siemens.com/press/en/pressrelease/?press=/en
/pressrelease/2011/fossil_power_generation/efp201105064.htm.
[7] Buck, R., E. Lpfert, and F. Tllez. Receiver for Solar-Hybrid Gas Turbine and CC
Systems (REFOS). Paper presented at the IEA Solar Thermal Conference, Sydney,
Australia, 2000.
[8] Heller, Peter, Markus Pfnder, Thorsten Denk, Felix Tellez, Antonio Valverde, Jess
Fernandez, and Arik Ring. Test and Evaluation of a Solar Powered Gas Turbine
System. Solar Energy 80, no. 10 (2006):1225-1230.
[9] Ho, Clifford K., Siri S. Khalsa, and Nathan P. Siegel. Modeling On-Sun Tests of a
Prototype Solid Particle Receiver for Concentrating Solar Power Processes and
Storage. Paper presented at the Proceedings of ASME Energy Sustainability, San
Francisco, CA, 2009.
[10] Z'Graggen, A., and A. Steinfeld. Heat and Mass Transfer Analysis of a Suspension of
Reacting Particles Subjected to Concentrated Solar Radiation - Application to the
Steam-Gasification of Carbonaceous Materials. International Journal of Heat and
Mass Transfer 52 (2009): 385-395.
[11] Hunt, A. J. A New Solar Thermal Receiver Utilizing Small Particles. Paper presented
at the Proceedings of the International Solar Energy Society Conference, Atlanta, GA,
May 1979.
[12] Abdelrahman, P., P. Fumeaux, and P. Suter. Study of Solid-Gas Suspension Used for
Direct Absorption of Concentrated Solar Radiation. Solar Energy 22, no. 1 (1979): 45-
48.
75

[13] Leary, P. L., and J. D. Hankins. A User's Guide for MIRVAL - A Computer Code for
Comparing Designs of Heliostat-Receiver Optics for Central Receiver Solar Power
Plants. Livermore, CA: Sandia Labs., 1979.
[14] Hunt, A. J., and C. T. Brown. Solar Testing of the Small Particle Heat Exchanger
(SPHER) #LBL-16497. Berkeley, CA: Lawrence Berkely National Laboratory, 1982.
[15] Bertocchi, R., J. Karni, and A. Kribus. Experimental Evaluation of a Non-Isothermal
High Temperature Solar Particle Receiver. Energy 29, no. 5-6 (2004): 687-700.
[16] Klein, Hanna Helena, Rachamim Rubin, and Jacob Karni. Experimental Evaluation of
Particle Consumption in a Particle Seeded Solar Receiver. Journal of Solar Energy
Engineering 130, no. 1 (2008).
[17] Miller, F. Radiative Heat Transfer in a Flowing Gas-Particle Mixture. PhD diss.,
University of California, Berkely, 1988.
[18] Klein, Hanna Helena, Jacob Karni, Rami Ben-Zvi, and Rudi Bertocchi. Heat Transfer
in a Directly Irradiated Solar Receiver/Reactor for Solid-Gas Reactions. Solar Energy
81, no. 10 (2007): 1227-1239.
[19] Rger, M., R. Buck, and H. Muller-Steinhagen. Numerical and Experimental
Investigation of a Multiple Air Jet Cooling System for Application in a Solar Thermal
Receiver. Journal of Heat Transfer 127, no. 8 (2005): 863-876.
[20] Karni, J., A. Kribus, B. Ostraich, and E. Kochavi, A High-Pressure Window for
Volumetric Solar Receivers. Journal of Solar Energy Engineering 120, no. 2 (1998):
101-107.
[21] Mande, Onkar Kiran. Window and Seal Design for a Small Particle Solar Receiver.
Masters Thesis, San Diego State University, 2012.
[22] Howell, John R., Robert Siegel, and M. Pinar Menguc. Thermal Radiation Heat
Transfer. New York: Taylor and Francis, 2002.
[23] Howell, J. R., and M. Perlmutter. Monte Carlo Solution of Thermal Transfer Through
Radiant Media Between Gray Walls. Journal of Heat Transfer 86, no. 1 (1964): 116-
122.
[24] Howell, J. R., and M. Perlmutter. Radiant Transfer Through a Gray Gas Between
Concentric Cylinders Using Monte Carlo. Journal of Heat Transfer 86, no. 2 (1964):
169-179.
[25] Huwaldt, Joseph A. Plot Digitizer. Sourceforge.net. Last modified November 03,
2012. http://plotdigitizer.sourceforge.net/.
[26] Bejan, A. Convection Heat Transfer. Hoboken, NJ: John Wiley & Sons, 2004.
[27] Kitzmiller, Kyle, and Fletcher Miller. Thermodynamic Cycles for Small Particle Heat
Exchange Receivers Used in Concentrating Solar Power Plants. Journal of Solar
Energy Engineering 133, no. 3 (2011): 031014. doi:10.1115/1.4004270.
[28] Incropera, Frank P., David P. Dewitt, Theodore L. Bergman, and Adrienne S. Lavine.
Introduction to Heat Transfer. 5th ed. Hoboken, NJ: John Wiley & Sons, 2007.
76

[29] Foundry Service & Supplies. Maxfire Blankets (1800F - 2600F Spun Insulation
Blanket). Blanket and Bulk Products. Accessed October 25, 2012.
http://supplies.foundryservice.com/viewitems/blanket-bulk-products/re-blankets-1800-
f-2600-f-spun-insulation-blanket-?&bc=100%7C1001%7C1020%7C1058.
[30] Kitzmiller, Kyle. Design, Construction, and Initial Testing of a Solar Simulator and
Lab-Scale Small Particle Solar Receiver. Masters Thesis, San Diego State University,
2012.
[31] Fluent Inc. FLUENT 6.3 User's Guide. ANSYS FLUENT. Accessed October 30,
2012. http://hpce.iitm.ac.in/website/Manuals/Fluent_6.3/fluent6.3/help/pdf/ug/pdf.htm.
77

APPENDIX

SUMMARY OF TURBULENCE MODEL


CONSTANTS USED IN FLUENT
78

All values are unchanged from the FLUENT defaults for the SST k- turbulence
model. For details on the function of these variables, refer to the FLUENT users manual
[31].
* 1
0.52
* 0.09
* 1.5
Mt0 0.25
a1 0.31
i(inner) 0.075
i(outer) 0.0828
TKE(inner) 1.176
TKE(outer) 1
SDR(inner) 2
SDR(outer) 1.168
EnergyPr 0.85
WallPr 0.85

Você também pode gostar