Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Modeling the small-scale dish-mounted solar

thermal Brayton cycle


Cite as: AIP Conference Proceedings 1734, 060002 (2016); https://doi.org/10.1063/1.4949144
Published Online: 31 May 2016

Willem G. Le Roux, and Josua P. Meyer

ARTICLES YOU MAY BE INTERESTED IN

Performance comparison of different thermodynamic cycles for an innovative central receiver


solar power plant
AIP Conference Proceedings 1850, 160024 (2017); https://doi.org/10.1063/1.4984558

Optical design and optimization of parabolic dish solar concentrator with a cavity hybrid
receiver
AIP Conference Proceedings 1734, 070002 (2016); https://doi.org/10.1063/1.4949149

A detailed radiation heat transfer study of a dish-Stirling receiver: The impact of cavity wall
radiation properties and cavity shapes
AIP Conference Proceedings 1734, 030017 (2016); https://doi.org/10.1063/1.4949069

AIP Conference Proceedings 1734, 060002 (2016); https://doi.org/10.1063/1.4949144 1734, 060002

© 2016 Author(s).
Modeling the Small-Scale Dish-Mounted Solar Thermal
Brayton Cycle
Willem G. Le Roux1, a) and Josua P. Meyer1, b)
1
Department of Mechanical and Aeronautical Engineering, University of Pretoria, South Africa
Private Bag X20, Hatfield, Pretoria 0028, South Africa
a)
Corresponding author: willem.leroux@up.ac.za
b)
josua.meyer@up.ac.za

Abstract. The small-scale dish-mounted solar thermal Brayton cycle (STBC) makes use of a sun-tracking dish reflector,
solar receiver, recuperator and micro-turbine to generate power in the range of 1-20 kW. The modeling of such a system,
using a turbocharger as micro-turbine, is required so that optimisation and further development of an experimental setup
can be done. As a validation, an analytical model of the small-scale STBC in Matlab, where the net power output is
determined from an exergy analysis, is compared with Flownex, an integrated systems CFD code. A 4.8 m diameter
parabolic dish with open-cavity tubular receiver and plate-type counterflow recuperator is considered, based on previous
work. A dish optical error of 10 mrad, a tracking error of 1⁰ and a receiver aperture area of 0.25 m x 0.25 m are
considered. Since the recuperator operates at a very high average temperature, the recuperator is modeled using an
updated ε-NTU method which takes heat loss to the environment into consideration. Compressor and turbine maps from
standard off-the-shelf Garrett turbochargers are used. The results show that for the calculation of the steady-state
temperatures and pressures, there is good comparison between the Matlab and Flownex results (within 8%) except for the
recuperator outlet temperature, which is due to the use of different ε-NTU methods. With the use of Matlab and Flownex,
it is shown that the small-scale open STBC with an existing off-the-shelf turbocharger could generate a positive net
power output with solar-to-mechanical efficiency of up to 12%, with much room for improvement.

INTRODUCTION
The open and direct STBC with recuperator is shown in Fig. 1a. The parabolic dish reflects and concentrates the
sun’s rays onto the receiver aperture so that the solar heat can be absorbed by the inner walls of the receiver. The
heat is transferred to air as working fluid, which makes this cycle attractive for water-scarce countries such as South
Africa. In the recuperator, hot turbine exhaust air is used to preheat compressed air before it enters the receiver. The
compressed and heated air expands in the turbine, which produces rotational shaft power for the compressor and the
electric load. The compressor, turbine and generator are mounted on a single shaft and all spin at the same rate [1,2].
It is simple, robust and easy to maintain. A recuperated STBC allows for lower compressor pressure ratios [3],
higher efficiency [4] and a less complex solar receiver, which operates at lower pressure.
The closed Brayton cycle was developed in the 1930s for power applications [4] and was adapted to the design
and development of STBCs for space power in the 1960s, with the success of lightweight and high-performance gas
turbines for aircraft. Experimental testing of the STBC has proven high reliability and efficiencies of between
11.76% [5] and above 30% [4], with turbine inlet temperatures of between 1 033 K and 1 144 K [4]. The Brayton
cycle is definitely worth studying when comparing its efficiency with other power cycles [6]. Micro-turbines can be
adapted from the small stationary gas turbine market, which allows for lower costs due to high production
quantities [7]. In South Africa, various micro-turbines are available off the shelf as turbochargers, for example, those
produced by Garrett [8]. The small-scale STBC has an advantage in terms of bulk manufacturability, mobility and
cost. The STBC can also be supplemented with natural gas as a hybrid system [9] for continuous operation. Storage
systems such as packed rock bed thermal storage [10] and lithium fluoride storage [11] can be coupled to the STBC.

SolarPACES 2015
AIP Conf. Proc. 1734, 060002-1–060002-8; doi: 10.1063/1.4949144
Published by AIP Publishing. 978-0-7354-1386-3/$30.00

060002-1
The hot exhaust air coming from the recuperator can be used to heat water or to run an absorption chiller. When this
heat is utilised, it makes the system very efficient and highly competitive.
Heat losses and pressure losses in the components of the cycle decrease the net power output of the system. A
solar receiver might be designed for high efficiency; however, when coupled to a Brayton cycle, it might not
perform well, due to it not being optimised to achieve a common goal together with other components. Optimisation
using the method of total entropy generation minimisation is considered to be a holistic optimisation approach and
has been applied to the STBC in recent work [12-15] for maximum net power output. Further modeling of the small-
scale STBC with turbocharger, for further development of an experimental setup at the University of Pretoria
(Fig. 1b), is considered in this paper. Recent studies are combined, updated, improved and applied so that results
from an analytical model of the small-scale STBC in Matlab can be compared with results from a Flownex model as
validation. A new SolTrace model and a new method to determine the recuperator effectiveness are included. The
amount of heat available for water heating is calculated and two different off-the-shelf turbochargers are considered.

Air out
11

10 9

Recuperator
3 4
8 W net  W t  W c
2

Compressor Turbine Load

Receiver
5 7
6

1 Air in Q *

a) b)
FIGURE 1. a) The open and direct STBC [12] where Q * is the available solar heat rate at receiver cavity. b) Experimental
setup showing the dish with diameter of 4.8 m and insulated solar receiver mounted on a two-axis solar tracking system [16].

OPEN-CAVITY TUBULAR RECEIVER


The open-cavity tubular solar receiver (see Fig. 1b and Fig. 2a) consists of a stainless steel tube which is coiled
to form a solar cavity. The receiver is covered with 100 mm thick ceramic fibre insulation. The receiver should be
able to operate at a maximum surface temperature of 1 200 K. Le Roux et al. [17] have shown the modeling and
optimisation of the receiver by assuming a solar tracking error of 1°, optical error of 10 mrad, dish reflectivity of
85% and a DNI of 1 000 W/m2. An aluminium parabolic solar dish with diameter of 4.8 m and rim angle of 45⁰ was
considered. According to [17], the optimum receiver aperture area is 0.25 m x 0.25 m. The factors contributing to
the temperature profile and net heat transfer rate on the receiver tube can be divided into two components:
geometry-dependent and temperature-dependent. The geometry-dependent factors include the concentrator dish with
its optics: tracking error, optical error, reflectance, spillage and shadowing. These factors were modeled with
SolTrace, a ray-tracing software. The optical error of the dish was determined from the slope error and specularity
error as shown in Eq. (1). A pillbox sunshape parameter of 4.65 mrad was assumed in the SolTrace analysis. The
reflectance of the receiver tube was assumed to be 15% (oxidised stainless steel). For the temperature-dependent
factors, a method to determine the receiver tube surface temperatures and net heat transfer rates along the length of
the receiver tube was described in [17], as shown in Eq. (2) - (3). It was shown that the open-cavity tubular solar
receiver surface temperature and net heat transfer rate for heating air depend on the receiver size, mass flow rate
through the receiver, receiver tube diameter, receiver inlet temperature and dish errors. For the specified dish and
receiver, receiver efficiencies of between 43% and 70% could be obtained with the open-cavity tubular receiver with

060002-2
mass flow rates of between 0.06 kg/s and 0.08 kg/s, tube diameters of between 0.05 m and 0.0889 m and air inlet
temperatures of between 900 K and 1 070 K [17].

 optical  4 slope y


1/ 2
2
  specularit
2
(1)

 n 1  Q
    1 1 
Q net , n   Ts ,n    net ,i   Tin , 0     (2)
  c
i 1  m
   hA 2 m
 c p0 
 p0    n 

   
N
Q net , n  Q * n  An  Fn  j  nTs4,n   j Ts4, j  An Fn    nTs4,n   j T4  hn An  An / R Ts , n  T  (3)
j 1

a) b)
FIGURE 2. a) The open-cavity tubular solar receiver model in SolTrace. b) Counterflow plate-type recuperator model [12].

METHODOLOGY

Analytical
For the solar receiver, the same receiver, dish and modeling method as was used in [17] is used for the analytical
model in this paper, except for the SolTrace model which is updated and improved as shown in Fig. 2a. The receiver
aperture area of 0.25 m x 0.25 m is considered. The counterflow plate-type recuperator model with channel length,
Lreg, and channel aspect ratio, a/b, are shown in Fig. 2b. The recuperator effectiveness is modeled using an updated
ε-NTU (effectiveness – number of transfer units) method [18]. This method takes the heat loss to the environment
into consideration when calculating the recuperator efficiency, since the recuperator operates at a very high average
temperature. The hot-side and cold-side efficiencies of the recuperator are calculated as shown in [19]. Compressor
and turbine maps from standard off-the-shelf turbochargers from Garrett [8] are used in the modeling to determine
the actual mass flow rate, pressure ratios and efficiencies. The turbine’s efficiency is determined with the blade
speed ratio which is a function of the inlet enthalpy, pressure ratio, turbine wheel diameter and rotational speed [15].
An exergy analysis was conducted for the system (see Fig. 1a) to determine the net power output, W net , as shown
in Eq. (4) [12-15]. Matlab was used as modeling tool. To maximise the net power output of the cycle, the objective
function, based on the method of total entropy generation minimisation, is presented in terms of the geometry
variables of the open-cavity tubular solar receiver and counterflow plate-type recuperator. The total internal entropy
generation rate, S gen ,int , is the sum of the entropy generation rate of each component in the system [15] and a
function of receiver and recuperator geometry variables. T* is the apparent sun’s temperature as an exergy source as
described in [15]. The compressor inlet temperature is assumed as 300 K and atmospheric pressure as 86 kPa. The
temperatures and pressures in the system are found with iteration using isentropic efficiencies of the compressor and
turbine as well as recuperator and receiver efficiencies. The maximum receiver surface temperature is constrained to

060002-3
1 200 K and the recuperator total plate mass is restricted to 500 kg. The maximum net power output of the system as
well as the optimum receiver and recuperator geometries are determined with a similar methodology as was used
in [19].

W net  T S gen ,int  1  T / T *Q *  m c p 0 T1  T11   m T c p 0 ln T1 / T11  (4)

FIGURE 3. Flownex model of the STBC.

Flownex
The results of the analytical modeling method are compared with results obtained with Flownex, an integrated
CFD code used for the design, simulation and optimisation of complete thermal fluid systems. Flownex has an easy-
to-use graphical user interface and results are presented in a graphical output. The complete integrated open STBC
with fixed components is modelled in Flownex at steady state by linking standard components together (Fig. 3). The
data input for the compressor and turbine maps of Flownex had to be converted from the data given on compressor
and turbine maps of Garrett [8]. Garrett-corrected-flow had to be converted to actual flow in kg/s which had to be
converted to Flownex-corrected-flow.

RESULTS
In SolTrace, a comparison was done of solar receiver power (final intersections of rays) by using the improved
SolTrace model (Fig. 2a) and the Soltrace model used in [17] which consisted of flat inner surfaces. An error of
about 3% was found when considering the dish optical error of 10 mrad and tracking error of 1⁰. This result shows
that receiver efficiencies from [17] could be used with negligible error in this paper. In Fig. 4, the maximum net
power output of the system at steady state is shown as a function of turbine pressure ratio, as was found with Matlab.
Each data point represents a maximum which is achieved by using a unique recuperator with specific dimensions as
shown in Table 1 for the turbocharger, GT2560R (compare with Fig. 4b). Table 1 shows that a recuperator with
a = 220 mm, b = 2.2 mm, L = 1.5 m and n = 45 is the best performing recuperator.
It is shown that for the 4.8 m diameter solar dish with 0.25 m x 0.25 m receiver aperture area, a large receiver
tube diameter allows for higher maximum net power output at high turbine pressure ratios. In previous work,
Le Roux et al. [17,19] found that a smaller receiver tube diameter allowed for higher first law efficiency but due to
the high pressure drop, a much lower second law efficiency was obtained, which indicated that a larger receiver tube
diameter would be more beneficial for the small-scale open STBC. The larger the receiver tube diameter, the smaller
the pressure drop through it and the more power is available at the turbine. According to [17], the highest second
law efficiencies for the receiver are achieved when the tube diameter and inlet temperature are large and the mass
flow rate small [17], creating a very high surface temperature. When the surface temperature is restricted to 1 200 K,
a larger mass flow rate and lower inlet temperature can still provide high second law efficiencies when a large tube
diameter is chosen. In Fig. 4, it is shown that as the turbine pressure ratio increases, the maximum net power output
increases up to a maximum. This effect is mostly due to a combination of compressor and turbine efficiencies which

060002-4
are a maximum when within a certain flow range and pressure ratio. Note that the curves are not smooth. This is due
to the efficiencies of the compressor and turbine which are sensitive to flow and pressure changes. Note that the
compressor efficiencies were interpolated from a compressor map. The recuperator efficiency influences the inlet
temperature to the receiver which influences receiver efficiency, maximum receiver surface temperature and turbine
power. The compressor power and compressor efficiency are in turn dependent on turbine power. The curves are
also not smooth because of the different recuperators used for each data point and due to the maximum receiver
surface temperature restriction.

2.5 1.2
D = 0.0833
D = 0.0625
2 1
D = 0.05 D = 0.0833

0.8 D = 0.0625
1.5

0.6
1
0.4

0.5
0.2

0 0
1.2 1.4 1.6 1.8 2 2.2 1 1.2 1.4 1.6 1.8

a) b)
FIGURE 4. Maximum net power output of the dish-mounted STBC using turbocharger (a) GT2052, (b) GT2560R.

TABLE 1. Optimum recuperator geometries, maximum net power output, maximum receiver
surface temperature and recuperator mass for GT2560R and receiver tube diameter, D = 0.0833 m.

rt a (m) b (mm) L (m) n W net (W) Ts,max (K) Mass (kg)


1.3125 0.15 4.5 2.2 22 419 1 199 250
1.375 0.37 4.5 2.2 15 737 1 196 410
1.4375 0.45 3.7 2.2 15 902 1 167 490
1.5 0.22 2.2 1.5 45 1 052 1 186 491
1.5625 0.22 2.2 1.5 45 1 098 1 130 491
1.625 0.22 2.2 1.5 45 959 1 076 491
1.6875 0.22 3 1.5 45 672 992 493
1.75 0.22 2.2 1.5 45 110 979 491

Note that for the GT2052 turbocharger, a maximum of 2.1 kW net power output was found (Fig 4a). This shows
that a solar-to-mechanical efficiency of 12% can be achieved with an off-the-shelf turbocharger. It should be noted
that a standard turbocharger was used, where turbine and compressor combination is already set and optimised for a
specific task in a vehicle. The net power output of the STBC can be significantly improved by investigating different
compressor and turbine combinations for maximum compressor and turbine efficiency.
Flownex was used as a tool for comparison of the results of the performance of the small-scale open STBC as
found analytically with Matlab. Case studies were carried out and results compared. A receiver tube diameter of
0.0833 m was used and the micro-turbine GT2560R was selected. Different steady-state operating speeds of the
micro-turbine were considered. The recuperator channel dimensions were chosen as a = 0.5 m, b = 2 mm, n = 40
and L = 0.6 m. The results as found with the Matlab model and with Flownex are shown in Table 2 for 87 000 rpm
and in Table 3 for 85 000 rpm with node reference to Fig. 1a. The temperature results in Table 2 are also shown in
Fig. 5. The results show that for the calculation of the steady-state temperatures and pressures, there is good
comparison between the Matlab results and Flownex results of within 8%, except for the recuperator outlet
temperature. Note that in Matlab, the ε-NTU efficiency was calculated with heat loss to the environment added,
while Flownex uses the simple ε-NTU method. The difference in recuperator outlet temperature can thus be
expected. Table 4 shows the different parameters found at steady state as calculated with Matlab and Flownex. Most
of the parameters compare well. The net power output however, does not compare well. Results also show that there
is a slight difference in the calculation between the shaft speed and mass flow rate. Further comparisons were done

060002-5
for GT2560R at other shaft speeds. When comparing results at 87 000 rpm in Flownex, it was also found that the
shaft speed was over-calculated in Matlab as 89 800 rpm and the mass flow rate was 0.0645 kg/s compared to
0.065 kg/s in Flownex. A possible cause is that compressor and turbine maps were inserted into Matlab and Flownex
in different ways by reading from the maps visually. These errors will contribute to differences in the calculation of
the shaft speed and mass flow rates as well as net power output. The net power output is extremely sensitive to shaft
speed and receiver inlet temperature. These sensitivities suggest that additional thermal storage or a gas burner
(hybrid system) will be very beneficial for the small-scale open STBC so that a constant inlet temperature can be
available for the turbine inlet at all times. Further work is required to compare the model with Flownex, using
different turbochargers at different operating points.

TABLE 2. Flownex and Matlab results for the system at 87 000 rpm (GT2560R).

Node Temperature (K) Pressure (kPa)


Difference
Flownex Matlab Flownex Matlab Difference
(%)
1 300.2 300.2 0 86 86 0
2 363.4 363.2 0.178 130.1 132.8 2.06
3 363.4 363.2 0.178 130.1 132.8 2.06
4 930.2 941.8 1.769 129.7 132.4 2.097
5 930.2 941.7 1.754 129.3 132.2 2.220
6 1 051.6 1 047 0.531 127.3 130.2 2.286
7 1 051.6 1 047 0.545 127.1 129.9 2.211
8 990.7 976.4 1.990 86.6 86.61 0.012
9 990.7 976.4 1.990 86.6 86.61 0.012
10 431.4 393.9 23.69 86 86 0

TABLE 3. Flownex and Matlab results for the system at 85 000 rpm (GT2560R).

Node Temperature (K) Pressure (kPa)


Difference
Flownex Matlab Flownex Matlab Difference
(%)
1 300.2 300.2 0 86 86 0
2 356.8 359.0 2.614 128 128.0 0.039
3 356.8 359.0 2.614 128 128.0 0.039
4 940.0 991.9 7.777 127.8 127.8 0.0156
5 940.0 991.8 7.763 127.5 127.6 0.0549
6 1 058 1 094 4.704 125.7 125.8 0.0636
7 1 058 1 094 4.684 125.5 125.5 0.00797
8 998.6 1 026 3.805 86.2 86.21 0.0116
9 998.6 1 026 3.805 86.2 86.21 0.0116
10 422.3 389.1 22.31 86 86 0

It should also be noted from Table 2 and Table 3 that a recuperator outlet temperature of close to 400 K is
available, which can be applied for water heating. For the GT2560R spinning at 85 000 rpm, a maximum water
heating capacity of 5.3 kW is available from the recuperator outlet according to Eq. (5). The energy utilisation factor
is the sum of the net power output and the rate of water heating divided by the available solar power on the dish. If
the water heating capacity is considered together with hybridisation, the cycle can be competitive in terms of
efficiency, cost and continuous operation. Further work is required on the modeling and optimisation of the water
heater and the ratio of electricity produced versus water heating capacity at the exhaust.

Q max,T111  m c p 0 T11  T1  (5)

060002-6
1200

1000

800

T (K)
600

400

200 Flownex
Matlab
0
1 2 3 4 5 6 7 8 9 10
Position in the cycle

FIGURE 5. Comparison between analytical and Flownex results using micro-turbine GT2560R at 87 000 rpm.

TABLE 4. Flownex and Matlab parameters with micro-turbine GT2560R at steady state.
Parameter Matlab Flownex
Reciever heat input (kW) 7.24 8.1
Shaft speed (rpm) 85 926 85 000
Mass flow rate (kg/s) 0.0593 0.060
Turbine isentropic efficiency (%) 61.93 62.2
Compressor isentropic efficiency (%) 61.3 63.4
Compressor pressure ratio 1.487 1.488
Turbine pressure ratio 1.456 1.456
Compressor power (kW) 3.663 3.568
Turbine power (kW) 4.690 4.109
Net power output (kW) 1.027 0.541
Recuperator efficiency (%) 95.5 91
Recuperator hot side h (W/m2K) 107 105

DISCUSSION AND CONCLUSION


The small-scale dish-mounted STBC with turbocharger was modeled in this work. Results were generated in
Matlab with the use of the method of total entropy generation minimisation and were compared with results from
Flownex as validation. A new SolTrace model and method to determine the recuperator effectiveness were also
considered in this work. The potential for water heating was also shown. For the calculation of the steady-state
temperatures and pressures, there was good comparison between the Matlab results and Flownex results (within 8%)
except for the recuperator outlet temperature, since different ε-NTU methods were used to calculate the recuperator
efficiency. Results showed that solar-to-mechanical efficiencies of up to 12% could be achieved when using a
standard off-the-shelf turbocharger. This efficiency is low compared to other solar technologies however; there are
many ways in which the efficiency can be improved. The current work only considered turbochargers with an
already-combined compressor and turbine. This combination is chosen for performance in a vehicle rather than for
power generation. Further compressor and turbine matching should be performed to identify better combinations of
these units for significant efficiency improvement. It should also be noted that a very simple solar receiver was used
for the modeling. The use of ceramic materials for higher receiver operating temperatures can also be considered.
The receiver efficiency can be improved by having more precise tracking and dish optics and higher dish
reflectivity. Note that in this paper, a receiver aperture area of 0.25 m x 0.25 m and a dish with poor tracking (1⁰
error) and dish optics (10 mrad error) were assumed. According to previous work [17], the receiver efficiency can be
increased with up to 30% when the optical error is changed to 5 mrad and the aperture area to 0.2 m x 0.2 m. The
receiver efficiency can be increased with up to 48% when the tracking error is changed to 0⁰ and the aperture area to
0.15 m x 0.15 m. Furthermore, this efficiency can be increased with up to 71% when both the optical error and
tracking error are changed to 5 mrad and 0⁰ respectively, with an aperture area of 0.1 m x 0.1 m. By adding water

060002-7
heating using the hot exhaust air, the overall efficiency will be increased further. It should also be noted that the
cycle offers continuous operation with the use of gas.
The validated STBC-models discussed in this paper will be valuable for the further experimental testing of this
cycle. Future work will include further increase in efficiency using the abovementioned points. With further
research, the small-scale open STBC could be a competitive small-scale solar energy solution to the people of South
Africa.

NOMENCLATURE
a Recuperator channel width (m) t Thickness (mm)
A Area (m) T Temperature (K)
b Recuperator channel height (m) W Power (W)
cp0 Constant pressure specific heat (J/kgK) ɛ Emmissivity
D Receiver tube diameter (m) σ Stefan-Boltzmann constant (W/m2K)
F View factor ω Error (mrad)
h Convection heat transfer coefficient (W/m2K) Subscripts
H Recuperator height (m) net Net
Lreg Recuperator channel length (m) c Compressor
m
 Mass flow rate (kg/s) t Turbine
n Number of recuperator flow channels in one direction s Surface
Q Heat rate (W) n Receiver tube section number along tube length
r Pressure ratio in At the inlet
R Thermal resistance, conduction (K/W)  Environment
S gen ,int Total internal entropy generation rate (W/K) max Maximum

REFERENCES
1. H. L. Willis and W. G. Scott, Distributed power generation (CRC Press, New York, 2000).
2. K. Shiraishi and Y. Ono, Mitsubishi Heavy Industries, Ltd. Technical Review 44(1) (2007).
3. B. S. Stine and R. W. Harrigan, Solar energy fundamentals and design (John Wiley & Sons, New York, 1985).
4. A. Pietsch and D. J. Brandes, “Advanced solar Brayton space power systems”, in Intersociety Energy
Conversion Engineering Conference, (Los Alamitos, CA, 1989), pp. 911-916.
5. B. Dickey, “Test results from a concentrated solar microturbine Brayton cycle integration”, in TurboExpo,
GT2011, ASME Conference Proceedings (ASME, Vancouver, British Columbia, Canada, 2011).
6. L. Chen, W. Zhang, F. Sun, Applied Energy 84, 512-525 (2007).
7. D. Mills, Solar Energy 76, 9-31 (2004).
8. Garrett, Garrett by Honeywell: turbochargers, intercoolers, upgrades, wastegates, blow-off valves, turbo-
tutorials. Available at: http://www.TurboByGarrett.com; 2009 [accessed 23.07.2015].
9. C.F. McDonald and C. Rodgers, Journal of Engineering for Gas Turbines and Power 124, 835-844 (2002).
10. K. G. Allen, “Performance characteristics of packed bed thermal energy storage for solar thermal power
plants”, Thesis, University of Stellenbosch, 2010.
11. H. M. Cameron, L. A. Mueller and D. Namkoong, Report NASA TM X-2552, Lewis Research Center, 1972.
12. W. G. Le Roux, T. Bello-Ochende and J. P. Meyer, Energy 36, 6027-6036 (2011).
13. W. G. Le Roux, T. Bello-Ochende and J. P. Meyer, Energy 46, 42-50 (2012).
14. W. G. Le Roux, T. Bello-Ochende and J. P. Meyer, Int. J. Energ. Res. 36, 1088-1104 (2012).
15. W. G. Le Roux, T. Bello-Ochende and J. P. Meyer, Renew. Sust. Energ. Rev. 28, 677–690 (2013).
16. W. G. Le Roux, T. Bello-Ochende and J. P. Meyer, “Experimental testing of a tubular cavity receiver for a
small-scale solar thermal Brayton cycle” in SASEC 2015, (Skukuza, Kruger National Park, South Africa).
17. W. G. Le Roux, T. Bello-Ochende and J. P. Meyer, Energ. Convers. Manage. 84, 457-470 (2014).
18. G. F. Nellis and J. M. Pfotenhauer, Journal of Heat Transfer 127, 1071-1073 (2005).
19. W. G. Le Roux, T. Bello-Ochende and J. P. Meyer, “Optimisation of an open rectangular cavity receiver and
recuperator used in a small-scale solar thermal Brayton cycle with thermal losses”, in HEFAT2014, (Orlando,
Florida, 2014).

060002-8

You might also like