Available online at ScienceDirect. Energy Procedia 49 (2014 ) SolarPACES 2013

Similar documents
Numerical Evaluation of the Extinction Coefficient of Honeycomb Solar Receivers

Available online at ScienceDirect. Energy Procedia 69 (2015 )

Characterization of high temperature solar thermal selective absorber coatings at operation temperature

Optical Properties, Surface and Geometry

Available online at ScienceDirect. Energy Procedia 69 (2015 )

Overview of solar receiver design

Radiative Equilibrium Models. Solar radiation reflected by the earth back to space. Solar radiation absorbed by the earth

Design of an innovative internal structure for surface solar air heater s

OPTICAL PERFORMANCE ANALYSIS OF A POROUS OPEN VOLUMETRIC AIR RECEIVER IN A SOLAR POWER TOWER

HYDRODYNAMICS AND RADIATION EXTINCTION CHARACTERISTICS FOR A FREE FALLING SOLID PARTICLE RECEIVER

Simplified Collector Performance Model

An on-site test method for optical efficiency of large-size parabolic trough collectors

COMPUTATIONAL FLUID DYNAMICS ANALYSIS OF PARABOLIC DISH TUBULAR CAVITY RECEIVER

The Peak Flux Constraints on Bladed Receiver Performance in High- Temperature Molten Salt Concentrating Solar Power Systems

INFLUENCE OF SURFACE EMISSIVITY AND OF LOW EMISSIVITY SHIELDS ON THE THERMAL PROPERTIES OF LOW DENSITY INSULATING MATERIALS

Indo-German Winter Academy

ScienceDirect. Heat transfer and fluid transport of supercritical CO 2 in enhanced geothermal system with local thermal non-equilibrium model

MAXIMUM NET POWER OUTPUT FROM AN INTEGRATED DESIGN OF A SMALL-SCALE OPEN AND DIRECT SOLAR THERMAL BRAYTON CYCLE. Willem Gabriel le Roux

The energy performance of an airflow window

2. Illustration of Atmospheric Greenhouse Effect with Simple Models

CFD Analysis of Solar Tower Hybrid Pressurized Air Receiver (HPAR) using a Dual-Banded Radiation Model

Available online at ScienceDirect. Energy Procedia 78 (2015 ) th International Building Physics Conference, IBPC 2015

Reading Problems , 15-33, 15-49, 15-50, 15-77, 15-79, 15-86, ,

Fig 1. Power Tower during Operation

Lecture 3a: Surface Energy Balance

Exergy Optimisation for Cascaded Thermal Storage

Lecture 3a: Surface Energy Balance

UNIT FOUR SOLAR COLLECTORS

THE COMBINED HEAT TRANSFER OF RADIATION AND MIXED CONVECTION ANALYSIS IN A LID-DRIVEN TRAPEZOIDAL CAVITY

Performance Assessment of PV/T Air Collector by Using CFD

Analysis of the Cooling Design in Electrical Transformer

Problem One Answer the following questions concerning fundamental radiative heat transfer. (2 points each) Part Question Your Answer

Absorptivity, Reflectivity, and Transmissivity

EXPERIMENTAL STUDY ON A CASCADED PCM STORAGE RECEIVER FOR PARABOLIC DISH COLLECTOR

PERFORMANCE EVALUATION OF REFLECTIVE COATINGS ON ROOFTOP UNITS

Postprint.

HEAT LOSS CHARACTERISTICS OF A ROOF INTEGRATED SOLAR MICRO-CONCENTRATING COLLECTOR

Radiation from planets

COMBINED MEASUREMENT OF THERMAL AND OPTICAL PROPERTIES OF RECEIVERS FOR PARABOLIC TROUGH COLLECTORS

Material and Design Requirements for Advanced Concentrators

Available online at ScienceDirect. Energy Procedia 49 (2014 ) SolarPACES 2013

This section develops numerically and analytically the geometric optimisation of

Mathematical Investigation and CFD Simulation of Monolith Reactors: Catalytic Combustion of Methane

ScienceDirect. System design of a 1 MW north-facing, solid particle receiver

Thermal Systems. What and How? Physical Mechanisms and Rate Equations Conservation of Energy Requirement Control Volume Surface Energy Balance

Radiation in the atmosphere

Available online at ScienceDirect. Energy Procedia 69 (2015 )

Description of radiation field

Radiation Heat Transfer. Introduction. Blackbody Radiation. Definitions ,

N. Lemcoff 1 and S.Wyatt 2. Rensselaer Polytechnic Institute Hartford. Alstom Power

exp ( κh/ cos θ) whereas as that of the diffuse source is never zero (expect as h ).

Outline. Stock Flow and temperature. Earth as a black body. Equation models for earth s temperature. Balancing earth s energy flows.

Chapter 1 INTRODUCTION AND BASIC CONCEPTS

Fluid Flow and Heat Transfer Characteristics in Helical Tubes Cooperating with Spiral Corrugation

NUMERICAL SIMULATIONS OF DIRECT ABSORPTION OF SOLAR RADIATION BY A LIQUID

Blackbody Radiation. A substance that absorbs all incident wavelengths completely is called a blackbody.

Radiation in climate models.

2. The Steady State and the Diffusion Equation

High Flux Solar Simulator and Solar Photo-Thermal Reactor: Characterization and Design

Photo-Thermal Engineering for Clean Energy and Water Applications

SOLVING TRANSIENT CONDUCTION AND RADIATION USING FINITE VOLUME METHOD

Beer-Lambert (cont.)

Total Thermal Diffusivity in a Porous Structure Measured By Full Field Liquid Crystal Thermography

Mechanism of Heat Transfer Enhancement in the Core Flow of a Tube and Its Numerical Simulation

1. The most important aspects of the quantum theory.

Heat Transfer: Physical Origins and Rate Equations. Chapter One Sections 1.1 and 1.2

Heat Loss from Cavity Receiver for Solar Micro- Concentrating Collector

Available online at ScienceDirect. Energy Procedia 74 (2015 )

Laser Beam Interactions with Solids In absorbing materials photons deposit energy hc λ. h λ. p =

Heat and Mass Transfer Unit-1 Conduction

Available online at ScienceDirect. Energy Procedia 78 (2015 ) th International Building Physics Conference, IBPC 2015

Lecture # 04 January 27, 2010, Wednesday Energy & Radiation

ANALYSIS OF HEAT TRANSFER IN TURBULENT FLOW THROUGH THE TUBE WITH UNIFORM HEAT FLUX

Heat Transfer. V2 4Jun15

Chapter 5: The First Law of Thermodynamics: Closed Systems

Radiative heat transfer

2. Stellar atmospheres: Structure

Thermal conversion of solar radiation. c =

PROBLEM L. (3) Noting that since the aperture emits diffusely, I e = E/π (see Eq ), and hence

Effect of Humidity on Thermal Performances of a Non-evacuated Parabolic Trough Solar Collector

Fundamentals on light scattering, absorption and thermal radiation, and its relation to the vector radiative transfer equation

Department of Energy Science & Engineering, IIT Bombay, Mumbai, India. *Corresponding author: Tel: ,

Lecture 28. Key words: Heat transfer, conduction, convection, radiation, furnace, heat transfer coefficient

Illumination, Radiometry, and a (Very Brief) Introduction to the Physics of Remote Sensing!

Thermal Radiation Studies for an Electron-Positron Annihilation Propulsion System

Radiative Transfer Multiple scattering: two stream approach 2

A model of the optical properties of a non-absorbing media with application to thermotropic materials for overheat protection

MNFFY 221/SIF 4082 Energy and Environmental Physics. 2002

Numerical Simulation of H 2 O/LiBr Falling Film Absorption Process

REPORT ON WORKSHOP: DEFINING RTE SOLUTION METHODS FOR MULTI-MODE TRANSFER. June 3-4, 2016, Istanbul Jack Howell and Pinar Mengüç

THERMAL AND OPTICAL CHARACTERIZATION OF PARABOLIC TROUGH RECEIVERS AT DLR S QUARZ CENTER RECENT ADVANCES

Effect of Periodic Variation of Sol-air Temperature on the Performance of Integrated Solar Collector Storage System

Spectrum of Radiation. Importance of Radiation Transfer. Radiation Intensity and Wavelength. Lecture 3: Atmospheric Radiative Transfer and Climate

atmospheric influences on insolation & the fate of solar radiation interaction of terrestrial radiation with atmospheric gases

Available online at ScienceDirect. Procedia Engineering 90 (2014 )

Div. 1 Div. 2 Div. 3 Div.4 8:30 am 9:30 pm 12:30 pm 3:30 pm Han Xu Ruan Pan

ME 476 Solar Energy UNIT TWO THERMAL RADIATION

Using LBM to Investigate the Effects of Solid-Porous Block in Channel

Lecture 3: Atmospheric Radiative Transfer and Climate

2 The Radiative Transfer Equation

Transcription:

Available online at www.sciencedirect.com ScienceDirect Energy Procedia 49 (2014 ) 408 417 SolarPACES 2013 Parametric study of volumetric absorber performance A. Kribus a, *, M. Grijnevich a, Y. Gray a and C. Caliot b a School of Mechanical Engineering, Tel Aviv University, Tel Aviv 69978, Israel b PROMES CNRS, 7 Rue du Four Solaire, 66120 Font-Romeu, France Abstract This study investigates the possible performance of volumetric absorbers as a function of geometric and material properties, aiming to identify the best absorber design parameters and the highest efficiency that may be expected. A simplified model is used with non-equilibrium heat transfer correlations and the two-flux approximation of one-dimensional radiative transfer. The radiative flux model has been validated against a detailed Monte-Carlo simulation. This model is simple enough for fast computation and parametric study. The results show that considerable gains in volumetric absorber efficiency can be achieved by careful selection of the absorber properties. In addition to porosity and pore size, two additional overlooked properties are also significant: the thermal conductivity of the absorber material, and the optical selectivity of the absorber surface. Under certain conditions, reducing the thermal conductivity leads to a significant decrease in emission loss, almost reaching the ideal volumetric absorber at local thermal equilibrium. This implies that the common preference for absorber materials such as SiC should be reconsidered. Spectral selectivity of the absorber material can also produce a significant increase in efficiency, but this is valid only when the selectivity is close to ideal. 2013 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/3.0/). 2013 The Authors. Published Elsevier Ltd. Selection and peer review by by the the scientific conference committee committee of SolarPACES of SolarPACES 2013 under 2013 responsibility under responsibility of PSE AG. of PSE AG. Final manuscript published as received without editorial corrections. Keywords: Volumetric effect, Porous medium, Solar gas turbine 1. Introduction Electricity generation from solar energy by thermo-mechanical conversion is currently limited in worldwide implementation. A major reason is the relatively low conversion efficiency (typically 15 18% annual average), which contributes to the high cost of the produced electricity, 2 3 times higher than electricity produced from * Corresponding author. Tel.: +972-3-6405924; e-mail address: kribus@tauex.tau.ac.il 1876-6102 2013 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/3.0/). Selection and peer review by the scientific conference committee of SolarPACES 2013 under responsibility of PSE AG. Final manuscript published as received without editorial corrections. doi:10.1016/j.egypro.2014.03.044

A. Kribus et al. / Energy Procedia 49 ( 2014 ) 408 417 409 conventional fossil fuels [1]. This level of performance and cost is achieved today in solar thermal power plant technologies (parabolic trough and power tower) that are based on steam cycles at moderate temperatures of 400 550 C. The solar dish-stirling technology is capable of achieving much higher efficiency, but it has not been able so far to demonstrate the reliability and cost that would make it a serious contender, and it is considered a niche solution for small distributed generation plants. A breakthrough in the competitiveness of utility-scale solar thermal electricity may occur if the conversion efficiency from sunlight to electricity can be significantly increased, and if the technology needed to achieve this high efficiency is reasonably simple, reliable and inexpensive. Nomenclature C p Specific heat of the fluid (J kg 1 K 1 ) Subscripts d p Typical pore size (m) a Aperture e b Blackbody emissive power (W m 2 ) f Fluid h Convection coefficient (W m 2 K 1 or W m 3 K 1 ) in Inlet i Radiation intensity (W m 2 sr 1 ) s Solid k Thermal conductivity of the absorber (W m 1 K 1 ) v Volumetric L Length (m) m Mass flow rate (kg s 1 ) Nu v Volumetric Nusselt number (=h v d 2 p k 1 f ) p Pressure (Pa) q R Radiative heat flux (W m 2 ) Re Reynolds number (=ρ u d p μ 1 ) T Temperature (K) u velocity (m s 1 ) Greek α v Volumetric absorption coefficient (m 1 ) β Extinction coefficient (m 1 ) ε p Absorber porosity ( ) ε a Front surface porosity ( ) λ c Critical wavelength (μm) ρ Density (kg m 3 ) μ Viscosity (kg m 1 s 1 ) Ω Albedo ( ) The desired breakthrough for solar thermal conversion may result from a technology for solar heating of air for high-temperature Combined Cycles (gas and steam turbines operating in series). Conventional Combined Cycles can reach high conversion efficiency of around 60% from heat to electricity, corresponding to overall solar to electricity conversion efficiency of 25 30% [2]. A Combined Cycle requires heating compressed air to temperatures above 1,000 C. Providing solar heat at these temperatures faces significant challenges, including materials for high temperature, optimization of radiative and convective heat transfer in the receiver, and optics for high concentration. Some major advances were achieved at the lab level for high-temperature receivers, for example heating air at 20 bar to 1,200 C [3]. However, most of the R&D work done in recent years focused on lower-temperature versions of airheating receivers, intended for indirect steam generation at air temperature of around 700 C [4], or for simple cycle gas turbine plants with solar air heating to about 800 C [5]. These lower temperature applications can be viewed as intermediate steps in the long-term vision, not yet reaching the highest possible efficiency, but providing valuable experience with solar air heating technologies. A key component in the solar thermal conversion process is the radiation absorber located at the focus of the concentrator field. For high temperature, a porous volumetric absorber should provide higher efficiency compared to a tubular receiver, due to the so-called volumetric effect [6]. However, volumetric absorbers tested to date do not show the expected effect: they produce high temperature at the front face and their efficiency is usually in the range

410 A. Kribus et al. / Energy Procedia 49 ( 2014 ) 408 417 70 80%, even for air temperature well below 1,000 C. For example the TSA receiver used a wire-knit volumetric absorber, with reported air exit temperatures up to 780 C and efficiency around 70% [7]. The ceramic grid HITREC absorber reported efficiency around 76% at 700 C, and 72% at 800 C [4]. The low efficiency may be attributed to the high temperatures found at the absorber front surface, similar to or even higher than the air outlet temperature [4][8], indicating that the volumetric effect was not achieved. This has also been confirmed in detailed simulations of volumetric absorbers [9], as shown in Fig. 1. Higher efficiency in the range 80 90% at air exit temperature of >1100 C were reported [3] but this was under very high incident radiation flux of several thousand suns, which is impractical for commercial plants. High efficiency was also reported with a high-temperature tubular absorber [10] but again under very high incident radiation flux. In addition to efficiency, the stability of the absorber materials under the high temperature and concentration and frequent thermal cycling is an issue that needs to be investigated. Another difficulty is the need to maintain air pressure: if the receiver is volumetric, then an expensive and possibly fragile high-pressure window is needed [11]. This work considers what is the maximum thermal efficiency that a volumetric absorber may reach under a reasonable incident radiation flux, and whether a significant volumetric effect can in fact be achieved. A significant volumetric effect is defined qualitatively here as a front surface temperature of the absorber that is significantly lower, i.e., by hundreds of degrees, than the fluid exit temperature. A previous theoretical estimate of the maximum efficiency of a volumetric absorber assumed local thermal equilibrium and presented results that are much higher than those achieved in practice [12]. A model with local thermal non-equilibrium is presented here, which is more realistic but still simple enough for fast computation and parametric study. This model is used to derive the influence of several geometric and material properties and find trends that may be applied in volumetric absorber design. Fig. 1. (a) Layout of the one-dimensional volumetric absorber model, (b) temperature profiles in a volumetric absorber showing high front surface temperature, from a detailed simulation [9] 2. Model A one-dimensional problem is assumed with changes occurring only in the axial direction perpendicular to the absorber surface, where the flow and radiation transport occur both along this direction, as shown in Fig. 1. The volumetric absorber is modeled as a homogeneous effective medium comprising two continuous phases, solid (porous absorber) and fluid (air). This is a common approach that constructs the model by averaging the transport equations on a representative elementary volume, which is larger than the pore scale but smaller than the problem scale [13]. This model is reasonable for structures such as ceramic foam but not for ordered structures such as a honeycomb. The solid and fluid phases have separate temperature distributions T s (x), T f (x), consistent with thermal non-equilibrium. The fluid mass flow rate is constant through each cross section due to the 1-D assumption and therefore the mass conservation equation is not needed. The momentum balance is represented with a Darcy- Forschheimer type correlation for the pressure drop in the porous medium along the flow in the axial direction [14]. The thermal energy balance is applied separately to the solid and fluid phases. For the fluid, the transport by conduction is neglected, assuming that it is significantly smaller than the transport by advection. The fluid is completely transparent to the radiation. For the solid, the area for conduction is determined by the porosity, and the

A. Kribus et al. / Energy Procedia 49 ( 2014 ) 408 417 411 interaction with the radiation (both incident from outside and emitted by the absorber) is included as the divergence of the radiative heat flux q R. dt mc f p dx = h T v s ( x) T f ( x) d 0 = ( 2 T 1 ε s p )k s dx h T 2 v s ( x) T f ( x) dq R dx (1) Inter-phase transport, or exchange of heat between the solid absorber and the air, in the heat balance equations is represented by an averaged convection coefficient per unit volume. The volumetric convection coefficient is derived from a correlation as a function of porosity and the Reynolds number based on the pore size [9]: Nu v = 32.504ε 0.38 p 109.94ε 1.38 p +166.65ε 2.38 3.38 ( p 86.98ε p )Re 0.438 (2) The volumetric absorber is modeled as a plane slab of a homogeneous participating medium for the radiative transport balance. The simple two-flux approximation is used to solve the radiative transport equation, in order to keep the computational effort to a minimum, even though it is known that this approach cannot represent well the directional distribution of the radiation in optically thick media. The approximation leads to two equations for the forward and backward intensities, i + and i : di + dx = 2βi + ( x)+ β Ω i + ( x)+ i ( x) di dx = 2βi ( x)+ β Ω i + ( x)+ i ( x) + 2β ( 1 Ω)e b ( x) π + 2β ( 1 Ω)e b ( x) π (3) The volumetric extinction coefficient and the albedo are found from correlations for porous media [9] as a function of the material s surface emissivity and the medium s porosity and pore size. Given the intensity, the net radiative flux is q R = π ( i + i ), and the net radiative heat source per unit volume absorbed by the medium that is needed for eq. (1) is derived from the divergence of the radiative flux: dq R dx = 2πα i + x v ( )+ i ( x) 4α e x v b ( ) (4) Eq. (3) and (4) refer to a gray medium. The treatment is extended to spectrally selective absorbers by defining two intensities in each direction to represent two spectral bands, separated by wavelength c. The set of radiative transport equations (3) is applied twice in the two spectral bands, where the optical properties are different in each band. The net radiative heat source then includes the contributions of the two spectral bands: dq R dx = 2πα i + v1 1 ( x)+ i 1 ( x) + 2πα i + v2 2 ( x)+ i 2 ( x) 4e x b ( ) (5) ( ) α v1 F λ c T s ( )+ α v2 1 F ( λ c T s ) Air mass flow rate per unit area is specified as a parameter rather than a boundary condition since the continuum equation is not solved. The thermal boundary conditions depend on an additional geometric parameter: the fraction of the inlet area that is free of the solid medium a, which can be the same as the bulk porosity p or different, depending on the method of preparation of the absorber. The frontal area of the solid medium, occupying an area fraction of (1 a ), is an important consideration in volumetric absorbers, since it receives the highest irradiance, and may cause local heating and reflection loss. The boundary condition for the solid phase accounting for the interaction of the solid frontal area with the radiation and with the fluid is:

412 A. Kribus et al. / Energy Procedia 49 ( 2014 ) 408 417 dt ( 1 ε a )α a q in = ( 1 ε a )α a σt 4 s ( 0) ( 1 ε s p )k s + 1 ε dx a 0 ( )h a T s ( 0) T in (6) The surface convection coefficient at the aperture h a is based on the analysis in [9]. The corresponding boundary condition for the fluid phase, including the convection of heat from the front surface before the fluid enters the bulk absorber, is: ( ) T f ( 0)= T in + 1 ε a h mc a T s ( 0) T in p (7) The radiation entering the bulk of the absorber, after a partial interception by the frontal area, is: i + ( 0)= ε a q in π (8) The back surface of the receiver is assumed to be perfectly insulated and a perfect diffuse reflector, so there are no losses through the back of the receiver. Due to the 1-D assumption in the model there are no losses via the sidewalls either. Therefore heat losses in this model include only radiative losses from the aperture surface, including reflection at the front surface, and emission and back scattering from inside the absorber. The set of equations and boundary conditions is solved using Matlab. The absorber efficiency is then calculated as the thermal power per unit area imparted to the fluid, divided by the incident flux: T f ( L) η = m C p ( T)dT q in (9) T in 3. Results 3.1. Validation The main uncertainty in the model is the use of the very simple two-flux approximation to represent the radiative transport in the absorber. The validity of the approximation was tested by comparison to a Monte-Carlo simulation and to a P1 approximation for a homogeneous medium with the same optical properties and with the same incident radiation and temperature profile. Fig. 2(a) shows the spatial distribution of the radiative heat source computed with the three models. The two-flux result is reasonably close to the more accurate MC result, while the P1 model that assumes high optical thickness has a much higher deviation. This supports the two-flux approximation as a reasonable model for radiative transport in the volumetric absorber. Fig. 2(b) shows another validation check, reproducing the solid and fluid temperature profile results of the detailed simulation reported in [9], for the same cases shown in Fig. 1(b) above. This case represents a SiC absorber with average pore diameter 1.5 mm and porosity 0.8, inlet air flow at atmospheric pressure and 300 K, and incident flux of 600 kw/m 2. The air inlet velocity was varied as shown in the figure. Clearly the results are very close even though the current model is simpler than the one used in [9].

A. Kribus et al. / Energy Procedia 49 ( 2014 ) 408 417 413 Fig. 2: (a) Comparison of two-flux approximation to MC and P1 simulations; (b) comparison to results of [9] as shown in Fig. 1(b) 3.2. Geometric parameters A set of definition and operation conditions common to all cases studied here is shown in Table 1. Typical absorber data, selected as a baseline case for the investigation of volumetric absorber performance, is shown as case A in Table 2. This corresponds to an open receiver where air enters at atmospheric pressure and ambient temperature. The material properties were selected to represent SiC foam, and were fixed at a constant value; a preliminary comparison of a temperature dependent thermal conductivity vs. a constant average value has shown that the average value is a good approximation. The temperature distribution across the absorber for case A is shown in Fig. 3 and numerical results are shown in Table 2. The solid temperature at the front surface is very close to the air exit temperature, and much higher than the local air temperature near the inlet. The absorber efficiency is then fairly low at 73.5%. The losses include emission and back scattering from the bulk of the absorber, and reflection of incident radiation at the front surface. The last two losses, which are independent on temperature, can be isolated by repeating the simulation with a high enough flow rate such that the temperatures are very low and emission is eliminated, leading to the maximum efficiency for each case as shown in Table 2. Note that the back scattering loss is high in the current work, since isotropic scattering was assumed for simplicity; this is not necessarily true for real absorbers [15]. For case A the isolation of the reflection and scattering effect leads to the conclusion that emission contributes a loss of 8.6%. Fig. 3: Temperature distributions for cases A (left) and B (right) Previous analyses have shown that higher porosity and smaller pore size tend to decrease the absorber front temperature [16], due to penetration of the radiation deeper into the absorber, and better convective heat transfer. Case B represents this improved absorber, as shown in Fig. 3 and Table 2. The absorber temperature near the front surface is reduced somewhat, but the volumetric effect is not significant: the absorber front temperature is still much

414 A. Kribus et al. / Energy Procedia 49 ( 2014 ) 408 417 higher than the local air temperature and close to the air exit temperature. The emission loss is reduced to 6.4%, and the absorber efficiency is improved to 74.7%. Table 1: Volumetric absorber data common to all cases Incident flux 600 kw m 2 Air inlet temperature Air inlet pressure 300 K 10 5 Pa Mass flow rate per unit area 0.6 kg s 1 m 2 Absorber thickness 0.010 m Absorber surface emissivity (gray) 0.9 Table 2: Variable parameters and performance results in case study Parameter A B C D E Porosity 0.8 0.9 0.8 0.9 0.9 Pore size (mm) 2 1 2 1 0.1 Absorber thermal conductivity (W m 1 K 1 ) 40 40 1 1 0.1 Air exit temperature T f(l) (K) 990 1001 985 1018 1027 Absorber front temperature T s(0) (K) 988 904 1114 814 564 Absorber efficiency 0.735 0.747 0.728 0.766 0.777 Maximum efficiency (scattering limit) 0.821 0.811 0.821 0.811 0.812 Front reflection loss 2% 1% 2% 1% 1% Back scattering loss 15.9% 17.9% 15.9% 17.9% 17.8% Emission loss 8.6% 6.4% 9.3% 4.5% 3.5% 3.3. Thermal conductivity A parameter that is often ignored or misunderstood is the thermal conductivity of the absorber. Cases C and D show the results for the previous two cases with a much lower thermal conductivity, representing an oxide ceramic instead of SiC (but assuming that the surface absorptivity remains at the high value). Fig. 4 shows the temperature distributions for the two cases with low thermal conductivity. For the case C with larger pores, the lower thermal conductivity produces higher front temperature, higher emission loss of 9.3%, and lower absorber efficiency. This is due to the power absorbed at the front surface, as defined in boundary condition (6), which cannot be effectively removed by convection in this case and causes a significant overheating of the front region. For case D with smaller pores, the convective heat transfer is higher, and the power absorbed at the front surface is removed more effectively by convection even when the conduction is low. The removal of the front surface effect reveals the main advantage of the low thermal conductivity: the ability of the solid absorber to maintain a temperature gradient along its depth as required for the volumetric effect. The low conductivity reduces heat conduction in the axial direction, and maintains the front region at lower temperature compared to the back of the absorber. Case D shows a closer resemblance to the ideal volumetric effect, leading to emission loss of only 4.5%, and the highest efficiency. Taking this trend to extreme, case E has pore size and thermal conductivity reduced by an order of magnitude from case D. This may be unrealistic for manufacturing, but the case is presented here to investigate the limit of performance when varying these parameters. Fig. 5 shows the temperature distribution in the absorber, which is almost identical to an ideal volumetric absorber, with the exception of a small temperature increase near the front surface due to the boundary condition. Nevertheless, the emission loss is still significant at 3.5% and the absorber efficiency is 77.7%, indicating that geometry and thermal conductivity have limited further improvement potential.

A. Kribus et al. / Energy Procedia 49 ( 2014 ) 408 417 415 Fig. 4: Temperature distributions for cases C (left) and D (right) Fig. 5: Temperature distributions for case E 3.4. Spectral selectivity Spectral selectivity is usually not considered at elevated temperatures, due to the increasing overlap between the solar spectrum and the emission spectrum. Nevertheless, it is possible to consider this possibility theoretically and evaluate its possible impact of the volumetric absorber performance. The simulations of cases D and E were repeated as cases DS, ES with an ideal selective absorber material having surface spectral emissivity ε λ ( λ )= 1 for λ<λ c and 0 for λ>λ c, where the cutoff wavelength λ c is a free value in the optimization for maximum efficiency. These results are of course theoretical limiting values, since no material is an ideal selective absorber, and selective properties are difficult to find at elevated temperatures. For properties that are slightly off the ideal case, the performance quickly deteriorates. For example, for spectral emissivity values of 0.9 and 0.1 instead of 1 and 0, the efficiency of case ES will reduce to 0.785, closer to the gray case E than to the ideal selective case ES. This is due to the porous nature of the absorber that acts similar to a blackbody cavity [17]: radiation emitted form the large surface area in the pores undergoes multiple reflections in the pore space, leading to higher overall emission from the front surface. The effective emissivity can then be much higher than the actual surface emissivity whenever the surface emissivity is non-zero. Table 3 presents the performance results and Fig. 6 shows the temperature distributions in these two cases. The emission losses in cases DS, ES are reduced to 0.7% and 0.4%, respectively, indicating that good volumetric performance is achieved with ideal spectral selectivity. The remaining loss is mostly due to back scattering (15.7%). These results are of course theoretical limiting values, since no material is an ideal selective absorber, and selective properties are difficult to find at elevated temperatures. For properties that are slightly off the ideal case, the performance quickly deteriorates. For example, for spectral emissivity values of 0.9 and 0.1 instead of 1 and 0,

416 A. Kribus et al. / Energy Procedia 49 ( 2014 ) 408 417 the efficiency of case ES will reduce to 0.785, closer to the gray case E than to the ideal selective case ES. This is due to the porous nature of the absorber that acts similar to a blackbody cavity [17]: radiation emitted form the large surface area in the pores undergoes multiple reflections in the pore space, leading to higher overall emission from the front surface. The effective emissivity can then be much higher than the actual surface emissivity whenever the surface emissivity is non-zero. Table 3: Performance results for cases DS and ES, the ideal selective surface version of cases D and E Parameter DS ES Porosity 0.9 0.9 Pore size (mm) 1 0.1 Absorber thermal conductivity (W m 1 K 1 ) 1 0.1 Cutoff wavelength (μm) 2.65 2.65 Air exit temperature T f(l) (K) 1078 1081 Absorber front temperature T s(0) (K) 871 587 Absorber efficiency 0.836 0.839 Maximum efficiency (scattering limit) 0.843 0.843 4. Discussion Fig. 6: Temperature distributions for cases D and E where the absorber material is an ideal selective surface The relatively simple model presented here can be used for fast simulation and evaluation of the influence of absorber design parameters, while detailed engineering design should add a more detailed simulation and experimental validation. The simple tool allows identification of interesting trends and questions that should be considered for future work. The approach of a volumetric absorber is supposed to produce a low emission loss via manipulation of the absorber temperature profile. It was shown that this desirable temperature profile is not easy to establish, and the absorber parameters such as porosity and pore size should be carefully optimized to produce the desired result. A new aspect of this issue is that when the geometric parameters are favorable, it is also necessary to reduce the thermal conductivity, in order to allow the existence of the temperature gradient without significant heat flow towards the aperture. Therefore, commonly preferred materials such as SiC may be actually detrimental to the goal of creating the volumetric temperature profile. When both geometric and thermal properties are set properly, then the loss by thermal emission can be reduced significantly, for example by a factor of more than 2 in the cases presented here.

A. Kribus et al. / Energy Procedia 49 ( 2014 ) 408 417 417 Another theoretical aspect shown here is that ideal spectral selectivity in a volumetric absorber can nearly eliminate the emission loss, even for cases with absorber temperature above 1,000 K. However, this does not lead to a practical design recommendation, since the advantage is quickly lost when the IR emissivity is non-zero. The main loss mechanism found in the simulations is back scattering within the absorber. This is partly due to a simplifying assumption of isotropic scattering in the model, but this mechanism should be considered carefully in real absorbers, and modeled with better accuracy. If this loss can be significantly reduced, for example, to half of the value calculated here, then the model predicts that absorber efficiencies close to 90% should be achievable. Acknowledgement This work was supported by a grant from the Ministry of Science and Technology, Israel, and by the Ministry of Foreign Affairs and the Ministry of National Education and Research, France. References [1] C. Philibert, Technology Roadmap: Concentrating Solar Power, International Energy Agency, 2010. [2] A. Kribus, R. Zaibel, D. Carey, A. Segal, J. Karni, A solar-driven Combined Cycle plant, Solar Energy. 62 (1998) 121 129. [3] A. Kribus, P. Doron, R. Rubin, R. Reuven, E. Taragan, S. Duchan, et al., Performance of the directly-irradiated annular pressurized receiver (DIAPR) operating at 20 bar and 1,200 C, Journal of Solar Energy Engineering. 123 (2001) 10 17. [4] B. Hoffschmidt, F.M. Tellez, A. Valverde, J. Fernandez, V. Fernandez, Performance Evaluation of the 200-kWth HiTRec-II Open Volumetric Air Receiver, Journal of Solar Energy Engineering. 125 (2003) 87 94. [5] P. Schwarzbözl, R. Buck, C. Sugarmen, A. Ring, M.J.M. Crespo, P. Altwegg, et al., Solar gas turbine systems: design, cost and perspectives, Solar Energy. 80 (2006) 1231 1240. [6] A.L. Ávila-Marín, Volumetric receivers in Solar Thermal Power Plants with Central Receiver System technology: A review, Solar Energy. 85 (2011) 891 910. [7] C. Tyner, G. Kolb, M. Prairie, Solar power tower development: recent experience, Sandia, 1996. [8] T. Fend, R.- Pitz-Paal, O. Reutter, J. Bauer, B. Hoffschmidt, Two novel high-porosity materials as volumetric receivers for concentrated solar radiation, Solar Energy Materials & Solar Cells. 84 (2004) 291 304. [9] Z. Wu, C. Caliot, G. Flamant, Z. Wang, Numerical simulation of convective heat transfer between air flow and ceramic foams to optimise volumetric solar air receiver performances, International Journal of Heat and Mass Transfer. 54 (2011) 1527 1537. [10] I. Hischier, Experimental and Numerical Analyses of a Pressurized Air Receiver for Solar-Driven Gas Turbines, Journal of Solar Energy Engineering. 134 (2012) 021003. [11] J. Karni, A. Kribus, B. Ostraich, E. Kochavi, A high-pressure window for volumetric solar receivers, Journal of Solar Energy Engineering- Transactions of the Asme. 120 (1998) 101 107. [12] W. Spirkl, H. Ries, A. Kribus, Optimal parallel flow in solar collectors for nonuniform irradiance, Journal of Solar Energy Engineering. 119 (1997) 156 159. [13] M. Quintard, S. Whitaker, Coupled, Nonlinear Mass Transfer and Heterogeneous Reaction in Porous Media, in: K. Vafai (Ed.), Handbook of Porous Media, 2nd ed., Taylor & Francis, Boca Raton, FL, 2005. [14] Z. Wu, C. Caliot, F. Bai, G. Flamant, Z. Wang, J. Zhang, et al., Experimental and numerical studies of the pressure drop in ceramic foams for volumetric solar receiver applications, Applied Energy. 87 (2010) 504 513. [15] J. Petrasch, P. Wyss, a. Steinfeld, Tomography-based Monte Carlo determination of radiative properties of reticulate porous ceramics, Journal of Quantitative Spectroscopy and Radiative Transfer. 105 (2007) 180 197. [16] Z. Wu, C. Caliot, G. Flamant, Z. Wang, Coupled radiation and flow modeling in ceramic foam volumetric solar air receivers, Solar Energy. 85 (2011) 2374 2385. [17]M.F. Modest, Radiative Heat Transfer, McGraw-Hill, New York, 1993.