Next Article in Journal
Optimal Day-Ahead Scheduling of a Smart Distribution Grid Considering Reactive Power Capability of Distributed Generation
Previous Article in Journal
One-Dimensional Modelling of Marine Current Turbine Runaway Behaviour
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Solar Hydrogen Production via a Samarium Oxide-Based Thermochemical Water Splitting Cycle

1
Department of Chemical Engineering, College of Engineering, Qatar University, P. O. Box 2713, Doha 2713, Qatar
2
Department of Chemical Engineering, Texas A&M University at Qatar, PO Box 23874, Doha 2713, Qatar
3
Department of Chemical and Biological Engineering, South Dakota School of Mines and Technology, Rapid City, SD 57701-3995, USA
4
Faculty of Engineering and Applied Science, University of Ontario Institute of Technology, 2000 Simcoe St. North, Oshawa, ON L1H 7K4, Canada
*
Author to whom correspondence should be addressed.
Energies 2016, 9(5), 316; https://doi.org/10.3390/en9050316
Submission received: 25 January 2016 / Revised: 17 April 2016 / Accepted: 18 April 2016 / Published: 25 April 2016

Abstract

:
The computational thermodynamic analysis of a samarium oxide-based two-step solar thermochemical water splitting cycle is reported. The analysis is performed using HSC chemistry software and databases. The first (solar-based) step drives the thermal reduction of Sm2O3 into Sm and O2. The second (non-solar) step corresponds to the production of H2 via a water splitting reaction and the oxidation of Sm to Sm2O3. The equilibrium thermodynamic compositions related to the thermal reduction and water splitting steps are determined. The effect of oxygen partial pressure in the inert flushing gas on the thermal reduction temperature (TH) is examined. An analysis based on the second law of thermodynamics is performed to determine the cycle efficiency (ηcycle) and solar-to-fuel energy conversion efficiency (ηsolar−to−fuel) attainable with and without heat recuperation. The results indicate that ηcycle and ηsolar−to−fuel both increase with decreasing TH, due to the reduction in oxygen partial pressure in the inert flushing gas. Furthermore, the recuperation of heat for the operation of the cycle significantly improves the solar reactor efficiency. For instance, in the case where TH = 2280 K, ηcycle = 24.4% and ηsolar−to−fuel = 29.5% (without heat recuperation), while ηcycle = 31.3% and ηsolar−to−fuel = 37.8% (with 40% heat recuperation).

Graphical Abstract

1. Introduction

Two concerns in many countries are long-term shortages of fossil fuels and the need to secure fuel supplies. Carbon dioxide emissions constitute another significant issue concerning all societies, as such emissions are considered the main cause of global warming and climate change, for which the consequences are predicted to be melting of glaciers, a sea level rise, and increased weather extremes. The CO2 emissions from fuel combustion were 32.1 Gt (billion metric tons) in 2015 [1]. As a consequence, it is important for countries to invest in and shift towards clean energy technologies such as solar energy.
The conversion of solar energy to chemical energy in the form of an alternative fuel such as H2 provides a promising future sustainable energy pathway. Hydrogen can be produced with solar thermal energy via water splitting reactions. Hydrogen used in this manner is an energy carrier. Its advantages include a high energy density (on a mass basis) and non-polluting nature.
The metal oxide (MO)-based two-step solar thermochemical cycle is a potentially advantageous way to produce hydrogen via water splitting [2,3,4]. The main reactions involved in the MO based two-step solar thermochemical water splitting cycle are as follows:
Reaction 1: Solar thermal dissociation of MO:
MO → MOreduced + O2
Reaction 2: Non-solar water splitting step:
MOreduced + H2O → MO + H2
The MO-based two-step solar thermochemical water splitting cycle is beneficial because it avoids the formation of an explosive gaseous mixture of H2 and O2 (since both gases are produced in different steps). Furthermore, the MO can be used in multiple thermochemical cycles as it is not consumed during any of the steps involved.
In the past, several MO-based redox systems were theoretically and experimentally studied as potential thermochemical water splitting reactions. These include zinc oxide cycles [2,3,4,5,6,7,8,9,10,11], iron oxide cycles [12,13,14,15], tin oxide cycles [16,17,18], terbium oxide cycles [19], mixed ferrite cycles [20,21,22,23,24,25], ceria cycles [26,27,28,29], and perovskite cycles [30,31,32]. In the case of zinc and tin oxide cycles, due to the volatile nature of the MOs involved, the loss of reactive material due to high temperature operation is inevitable. In the case of the iron oxide, ferrite, ceria, and perovskite cycles, the complete reduction of the MOs is not possible and hence a lower H2 production is expected (per kg of metal oxide and per unit solar energy input). Therefore, investigations of new thermochemical cycles aimed at H2 production via solar thermochemical water splitting are needed.
In this paper, the thermodynamic feasibility of a novel samarium oxide-based two-step solar thermochemical water splitting (Sm-WS) cycle is investigated. This is, to the best of the authors’ knowledge, the first time that a Sm-WS cycle has been investigated for use in solar H2 production. Thermodynamic modeling of the Sm-WS cycle is performed (in two sections) using the commercial thermodynamic HSC Chemistry software and databases [33]. In section one, the thermodynamic equilibrium compositions during the solar thermal reduction of Sm2O3 and the oxidation of Sm via water splitting are determined. The effect of the oxygen partial pressure in the inert flushing gas used inside a solar reactor during thermal reduction of Sm2O3 is examined. Also, the variation in the Gibbs free energy as a function of water splitting temperature (TL) for the H2 production step is estimated. In section two, energy and exergy analyses of the Sm-WS cycle are carried out. The solar reactor absorption efficiency, the required solar energy input, radiation heat losses, the rate of heat rejected by the quench unit and the water splitting reactor, and the solar-to-fuel conversion efficiency of the Sm-WS cycle are determined via computational thermodynamic modeling and the results are reported. A typical Sm-WS cycle is presented in Figure 1.
To perform the thermodynamic calculations in a conventional way, we rely on the experimental or assessed data (which utilizes stability functions) published in thermodynamic books and scientific journals. This approach is time consuming due to the difficulty in locating the thermodynamic data in the published literature and also because the calculations involved are complicated. Furthermore, inconsistencies arising from the use of different standard and reference states make the final results uncomparable. HSC Chemistry software and its databases permit rapid and straightforward conventional thermodynamic calculations on personal computers. The software provides calculation procedures for investigating the influence of different variables on a chemical system at equilibrium. In addition to the thermodynamic calculations, with the help of HSC Chemistry software and its databases, heat and material balance calculations for processes can be carried out straightforwardly, especially compared to manual methods. The dissolution and corrosion behavior of materials can also be studied in an expedite way using the Eh-pH-diagrams available in HSC Chemistry software. One of the limitations associated with the software is its inability to solve all chemical problems as it does not take into account the kinetics of chemical reactions. Nevertheless, it can be used to determine the optimal reaction conditions and yields of experiments (without carrying out actual expensive experiments) in an economical and effective manner. The name of the software is based on the fact that it performs calculations utilizing thermochemical database related to enthalpy (H), entropy (S), and heat capacity (Cp), for more than 28,000 chemical species. Due to these features, HSC has been widely used for a range of applications in academia, industry and research.
With an average concentration of about 8 parts per million (ppm), samarium is the 40th most abundant element in the Earth's crust. It is the fifth most abundant lanthanide and is more common than elements such as tin. Samarium oxides are readily available from commercial suppliers such as Sigma Aldrich. It can be synthesized inexpensively in laboratories using various synthesis approaches, such as combustion synthesis, the sol-gel method and the co-precipitation method, as samarium precursors such as samarium nitrate are available at a low cost. Samarium does not pose a threat to environment, including plants and animals. The melting points for Sm2O3 and Sm are 2608 and 1347 K, respectively. Similarly, the Δ H and Δ S values for the decomposition of Sm2O3 (according the reaction chemistry mentioned in Figure 1) are equal to 2240.5 kJ and 0.527 kJ/K, respectively.

2. Chemical Thermodynamic Equilibrium

The Sm-WS cycle mainly involves the following reactions:
Sm 2 O 3 2 Sm +   1.5 O 2 ( g )
2 Sm + 3 H 2 O Sm 2 O 3 +   3 H 2 ( g )
According to the reaction chemistry, three moles of H2 can be produced in one complete thermochemical cycle including one thermal reduction and one water splitting step. This seems to be an advantage of the Sm-WS cycle over previously investigated thermochemical cycles, as they are capable of producing either 1 or less than 1 mole of H2 in one complete thermochemical cycle. To perform the energy and exergy analyses, thermodynamic data for all the reactive species involved (as shown in Equations (3) and (4)) are taken from HSC and the analyses are performed assuming continuous operation of the solar reactor with an inlet molar flow rate of Sm 2 O 3 of 1 mol·s−1.
As reported in previous studies [14,19,29], the temperature required for the thermal reduction of MOs can be reduced using an inert flushing gas such as ultra-high purity Ar with an oxygen partial pressure in the range of 10−5 to 10−8 bar. From this viewpoint, the effect of the oxygen partial pressure in the inert flushing gas on the thermal reduction of Sm2O3 is examined. According to Figure 2, as the oxygen partial pressure in the inert flushing gas is decreased, TH can be significantly reduced. For instance, at oxygen partial pressure of 10−5 bar, the complete thermal reduction of Sm2O3 can be achieved at 3000 K. As the oxygen partial pressure is further lowered to 10−6, 10−7, and 10−8 bar, the required TH values decease to 2780, 2540, and 2280 K, respectively.
The influence of oxygen partial pressure in the inert flushing gas on the equilibrium compositions associated with the thermal reduction of Sm2O3 is also studied. As per the results in Figure 3a–c, the slope of the decrease in the molar concentration of Sm2O3 and the increase in the molar concentration of Sm(g) and O2(g) both shift towards lower TH with a reduction in the oxygen partial pressure in the inert flushing gas. Because of the lower value of oxygen partial pressure in the inert flushing gas, the entropy of the product gases increases and hence thermal reduction is possible at lower temperatures.
The variation in the Gibbs free energy associated with the water splitting reaction (Figure 4) indicates that H2 generation via water splitting using Sm produced after the reduction of Sm2O3 is possible below 6800 K (for a pressure of 1 bar). For a decrease in TL from 6800 to 1900 K, Δ G w s reduces by 1010 kJ·mol−1. However, further lowering of TL from 1900 to 300 K does not affect Δ G w s significantly.

3. Energy and Exergy Analyses

An exergy analysis of the Sm-WS cycle is carried out to provide a second law perspective. The process flow diagram for the Sm-WS cycle is shown in Figure 5. This cycle is comprised of a solar reactor, a quench unit, water splitting reactor, and a theoretical H2/O2 fuel cell. To perform the exergy analysis, the following assumptions are made:
  • The production of H2 is carried out at 1 bar and steady state conditions, with negligible viscous losses and changes in kinetic and potential energies.
  • The solar reactor is considered a perfectly insulated blackbody absorber.
  • The effective emissivity and absorptivity is both equal to unity.
  • There are negligible convective/conductive heat losses.
  • All reactions reach 100% completion.
  • All products separate naturally without expending any work.
  • Heat exchangers required for recovering the sensible and latent heat are omitted.
The exergy analysis is performed following the methodology and governing equations derived and used previously for other MO based solar thermochemical cycles [6,9,10]. Thermodynamic properties are extracted from HSC software and databases and the analysis is normalized to a unit Sm2O3 molar flow rate (1 mol·s−1) entering the solar reactor.
As a first step in the exergy analysis, the absorption efficiency ( η a b s o r p t i o n ), which is defined as the net rate at which energy is absorbed by the solar reactor divided by the solar energy input through the aperture, is determined as follows:
η a b s o r p t i o n =   1 ( σ T H 4 I C )
Here, I denotes the direct normal solar irradiance i.e., normal beam insolation (W/m), C denotes the solar flux concentration ratio (suns), TH denotes the solar reactor temperature required for the thermal reduction of Sm2O3 and σ denotes the Stefan-Boltzmann constant ( 5.6705 × 10 8 W·m−2·K−4).
Values for η a b s o r p t i o n calculated as a function of TH are presented in Figure 6a. The trends indicate that η a b s o r p t i o n increases with decreasing TH. For instance, at TH = 3000 K (oxygen partial pressure in the inert flushing gas = 10−5 bar), η a b s o r p t i o n is 54%. As oxygen partial pressure in the inert flushing gas is further reduced to 10−8 bar, the corresponding TH decreases by 720 K and η a b s o r p t i o n rises to 84.7%.
The influence of C on η a b s o r p t i o n (at constant TH = 2280 K and oxygen partial pressure = 10−8 bar in the inert flushing gas) is also examined. At C = 2000 suns, η a b s o r p t i o n is 23.4%. The efficiency η a b s o r p t i o n can be increased to 38.3, 51.1, 57.5, and 61.3% by raising C to 4000, 6000, 8000, and 10,000 suns, respectively (Figure 6b).
To operate the Sm-WS cycle, solar energy is required 1) to heat the Sm2O3 from TL to TH and 2) for complete thermal reduction of Sm2O3 as per Equation (3). For comparison purposes, the heating of Ar is not included in this calculation as it was not considered in previous studies. The net solar energy absorbed by the solar reactor to perform the above mentioned tasks can be calculated as:
Q r e a c t o r n e t = Q S m 2 O 3 h e a t i n g + Q S m 2 O 3 r e d u c t i o n
Q S m 2 O 3 h e a t i n g = n ˙ Δ H | S m 2 O 3 @ T L S m 2 O 3 @ T H
Q S m 2 O 3 r e d u c t i o n = n ˙ Δ H | Sm 2 O 3 @ T H 2 Sm ( g ) + 1.5 O 2 ( g ) @ T H
As per Figure 7, at TH = 3000 K (when oxygen partial pressure = 10−5 bar in the inert flushing gas), Q r e a c t o r n e t is 2583 kW. With a further decrease in oxygen partial pressure in the inert flushing gas, TH reduces and the corresponding Q r e a c t o r n e t values also decline. The value of Q r e a c t o r n e t decreases by 82.3 and 120 kW with a decline in TH to 2540 and 2280 K respectively as the oxygen partial pressure in the inert flushing gas is lowered to 10−7 and 10−8 bar.
Based on the calculated values for η a b s o r p t i o n and Q r e a c t o r n e t , the solar energy input required for the operation of Sm-WS cycle can be calculated as follows:
Q s o l a r =   Q r e a c t o r n e t   η a b s o r p t i o n
The variation in Q s o l a r as a function of TH is shown in Figure 7. It is seen that the maximum value of Q s o l a r (4777 kW) is needed at TH = 3000 K (oxygen partial pressure in the inert flushing gas = 10−5 bar). Furthermore, the required value of Q s o l a r decreases as TH reduces. For instance, at TH = 2780 K (oxygen partial pressure = 10−6 bar in the inert flushing gas), Q s o l a r is 3248 kW. An additional drop in TH to 2540 K (oxygen partial pressure = 10−7 bar in the inert flushing gas) reduces Q s o l a r to 3273 kW. The reason behind this lowering of Q s o l a r is the increase in η a b s o r p t i o n due to the decline in TH. Overall, as the oxygen partial pressure in the inert flushing gas decreases from 10−5 to 10−8 bar, TH declines by 720 K and Q s o l a r by 39.1%.
Due to operation at elevated temperatures, radiation heat losses from the solar reactor are inevitable. The radiation heat losses associated with the solar reactor installed for the Sm-WS cycle can be calculated as:
Q r e r a d i a t i o n = Q s o l a r Q r e a c t o r n e t
The values of Q r e r a d i a t i o n reported in Table 1 indicate that radiation heat losses are greatest (2194 kW) when TH = 3000 K (oxygen partial pressure in the inert flushing gas = 10−5 bar). A decrease in TH results in a significant reduction in Q r e r a d i a t i o n . This is again due to the fact that η a b s o r p t i o n is higher at lower values of TH. The quantity Q r e r a d i a t i o n can be lowered by 1748 kW with a decrease in TH to 2280 K (oxygen partial pressure = 10−8 bar in the inert flushing gas), which is a very significant reduction in input heat.
Products of the thermal reduction of Sm2O3 (i.e., Sm and O2) leaving the solar reactor are in the gas phase and also at high TH. At such TH levels, these gaseous products have a tendency to recombine and convert to Sm2O3. To avoid the re-formation of Sm2O3, the gaseous products exiting the solar reactor at TH are rapidly cooled to TL using a quench unit. During quenching, it is assumed that the chemical compositions of the products remain unaltered. The latent and sensible heat rejected by the quench unit can be determined as follows:
Q q u e n c h = n ˙ Δ H | 2 Sm ( g ) + 1.5 O 2 ( g ) @ T H 2 Sm ( s ) + 1.5 O 2 ( g ) @ T L
The thermal energy required for the separation of O2 and the inert Ar (gas separation) is not considered for comparison purposes, as it was excluded in previous studies.
Table 1 reports the change in Q q u e n c h with respect to TH. It can be seen that, at higher TH, Q q u e n c h is also elevated. For instance, at TH = 3000 K (oxygen partial pressure in the inert flushing gas = 10−5 bar), Q q u e n c h is 756 kW. The data also confirm that the heat loss due to quenching can be reduced if the solar reactor is operated at a lower oxygen partial pressure in the inert flushing gas, which also results in a lower TH. For instance, at an oxygen partial pressure in the inert flushing gas of 10−8 bar, the required values are TH = 2280 K and Q q u e n c h = 636 kW.
Irreversibilities associated with the solar reactor and quench unit due to the non-reversible chemical transformations and re-radiation losses to the surroundings can be calculated as:
I r r r e a c t o r = ( Q s o l a r T H ) + ( Q r e r a d i a t i o n 298 ) + n ˙ Δ S | Sm 2 O 3 @ T L 2 Sm ( g ) + 1.5 O 2 ( g ) @ T H
I r r q u e n c h = ( Q q u e n c h 298 ) + n ˙ Δ S | 2 Sm ( g ) + 1.5 O 2 ( g ) @ T H 2 Sm ( s ) + 1.5 O 2 ( g ) @ T L
Similarly, Q s o l a r and Q r e a c t o r n e t , as well as I r r r e a c t o r and I r r q u e n c h , are also observed to be greatest at TH = 3000 K (oxygen partial pressure in the inert flushing gas = 10−5 bar). As shown in Figure 8, the HSC simulation results indicate an 85.1% decrease in I r r r e a c t o r for a reduction in TH from 3000 to 2280 K as the oxygen partial pressure in the inert flushing gas decreases to 10−8 bar. Correspondingly, I r r q u e n c h is also lowered by 0.5 kW·K−1 due to a similar decline in TH. These trends are in good agreement with the results reported in previous studies [9,10].
Solid Sm obtained after the quench unit (at 298 K) is transferred to the water splitting reactor (Sm oxidizer) and reacted with water at TL = 298 K producing solar H2. The rate of heat rejected to the surroundings by the water splitting reactor is estimated as 970 kW using the following equation and assuming 100% conversion:
Q S m   o x i d a t i o n = n ˙ Δ H | 2 Sm + 3 H 2 O Sm 2 O 3 +   3 H 2 ( g ) @ T L
Also, the irreversibility associated with the water splitting reactor is observed to be 3.5 kW·K−1 using the following equation:
I r r S m   o x i d a t i o n = (   Q S m   o x i d a t i o n 298 ) + n ˙ Δ S | 2 Sm + 3 H 2 O Sm 2 O 3 +   3 H 2 ( g ) @ T L
An ideal H2/O2 fuel cell is added to the Sm-WS cycle to extract the maximum work from the net H2 produced. In this study, the efficiency of the H2/O2 fuel cell is considered as 100% and hence it is defined as an ideal fuel cell. The following two equations can be used to determine the theoretical rate of work performed (711 kW) and heat energy released (146 kW) by the H2/O2 fuel cell:
W F C I d e a l = n   ˙ Δ G | 3 H 2 ( g ) + 1.5 O 2 ( g ) 3 H 2 O ( l ) @ 298 K
Q F C I d e a l = ( 298 ) × n   ˙ Δ S | 3 H 2 ( g ) + 1.5 O 2 ( g ) 3 H 2 O ( l ) @ 298 K
In the case of solar thermochemical cycles, the cycle efficiency ( η c y c l e ) is defined as the ratio of theoretical work performed by the ideal fuel cell to the solar energy input:
η c y c l e =     W F C I d e a l Q s o l a r
Similarly, the solar-to-fuel-conversion efficiency (   η s o l a r t o f u e l ) can be defined as the ratio of higher heating value (HHV) of the H2 produced to the solar energy input:
η s o l a r t o f u e l =   H H V H 2 Q s o l a r
Both η c y c l e and η s o l a r t o f u e l of the Sm-WS cycle are estimated based on the HSC simulation results and given in Table 2. The values for η c y c l e and η s o l a r t o f u e l increase as TH decreases. At TH = 3000 K (oxygen partial pressure in the inert flushing gas = 10−5 bar), η c y c l e is 14.9% and η s o l a r t o f u e l is 17.9%. For a reduction in TH to 2280 K due to the decline in oxygen partial pressure in the inert flushing gas to 10−8 bar, both η c y c l e and η s o l a r t o f u e l can be increased up to 24.5% and 29.5%, respectively. These efficiency values are comparable with other MO cycles such as the ZnO/Zn cycle (29%) [5], the SnO2/SnO cycle (29.8%) [33], the Fe3O4/FeO cycle (30%) [34] and the ceria cycle (20.2%) [29], which all were investigated previously employing similar operating conditions.
By applying heat recuperation, η c y c l e and η s o l a r t o f u e l for the Sm-WS cycle can be further enhanced [19,29,35]. Heat rejected by the quench unit and the water splitting reactor can be recycled and re-utilized to drive the Sm-WS cycle. The total amount of heat that can be recuperated is calculated as:
Q r e c u p e r a b l e = Q q u e n c h +   Q S m   o x i d a t i o n
With the inclusion of the heat recuperation, the solar energy required for the operation of the cycle can be reduced as follows:
Q s o l a r ,   w i t h   r e c u p e r a t i o n = Q s o l a r [ ( %   r e c u p e r a t i o n ) × Q r e c u p e r a b l e ]
As an example, the alteration in Q s o l a r ,   w i t h   r e c u p e r a t i o n and Q r e c u p e r a b l e as a function of % heat recuperation for the case of TH = 2280 K (oxygen partial pressure in the inert flushing gas = 10−8 bar) is shown in Figure 9. The results indicate that, as the % heat recuperation increases, Q r e c u p e r a b l e rises whereas Q s o l a r ,   w i t h   r e c u p e r a t i o n declines. For instance, with no heat recuperation, Q s o l a r ,   w i t h   r e c u p e r a t i o n = 2909 kW. But as heat recuperation increases to 50%, Q s o l a r ,   w i t h   r e c u p e r a t i o n decreases to 2106 kW, with 803 kW of the thermal energy recycled from the quench unit and water splitting reactor to drive the Sm-WS cycle.
The efficiencies η c y c l e and η s o l a r t o f u e l after heat recuperation can be determined as:
η c y c l e =     W F C I d e a l Q s o l a r , w i t h   r e c u p e r a t i o n
  η s o l a r t o f u e l =   H H V H 2 Q s o l a r , w i t h   r e c u p e r a t i o n
The effects of heat recuperation on η c y c l e and η s o l a r t o f u e l for all TH values are provided in Table 3. Also, the increase in η c y c l e and η s o l a r t o f u e l due to heat recuperation for the case of TH = 2280 K (oxygen partial pressure in the inert flushing gas = 10−8 bar) is presented in Figure 10. The data reported in Table 3 and Figure 10 confirm that, as heat recuperation is employed, both η c y c l e and η s o l a r t o f u e l increase significantly. For example, at TH = 2280 K, η c y c l e and η s o l a r t o f u e l (without heat recuperation) are 24.5% and 29.5%. However, if 60% heat recuperation is employed, both η c y c l e and η s o l a r t o f u e l can be raised further, to 36.5% and 44.1%, respectively. These efficiency values are much higher than for previously examined MO cycles (Table 4).
The exergy analysis performed for the Sm-WS cycle can be verified by conducting an energy balance and evaluating the maximum achievable efficiency from the total available work and from the total solar power input. The energy balance (for all TH) confirms that:
W F C I d e a l = Q s o l a r ( Q r e r a d i a t i o n + Q q u e n c h +   Q S m   o x i d a t i o n + Q F C I d e a l )
At TH = 2540 K, the energy balance in Equation (24) indicates that W F C I d e a l   = 712 kW which is equal to W F C I d e a l determined by Equation (16). Likewise, at TH = 2280 K, the energy balance indicates that W F C I d e a l = 712 kW which is again equal to the value of W F C I d e a l determined by Equation (16). The available work can be estimated as the addition of the work done by the fuel cell and the work lost due to the irreversibilities in the solar reactor, quench unit, and water splitting reactor. In case of the Sm-WS cycle, the maximum cycle efficiency can be calculated as:
η c y c l e , m a x i m u m = W F C I d e a l + T L ( I r r r e a c t o r + I r r q u e n c h + I r r s m   o x i d a t i o n ) Q s o l a r
For all values of TH, it is observed that η c y c l e , m a x i m u m is equal to the efficiency of a Carnot heat engine operating between hot and cold temperature reservoirs:
η c y c l e , m a x i m u m = 1 T L T H =   η c a r n o t
For instance, at TH = 2280 K and TL = 298 K, η c y c l e , m a x i m u m = 86.9% which is equal to η c a r n o t = 86.9%. Similarly, at TH = 2780 K and TL = 298 K, η c y c l e , m a x i m u m = 89.3% which is again equal to η c a r n o t = 89.3%.
From the results obtained during this investigation (phase–1: thermodynamic analysis), we have realized that the temperatures required in case of Sm-WS cycle are significantly higher. To overcome this limitation, we are currently working towards improving the redox reactivity of Sm2O3 at lower temperatures by synthesizing doped Sm2O3. The host, i.e., Sm2O3, will be doped by suitable metal cations such as Ni, Mn, Co, Fe, Ce, and others using various synthesis approaches such as sol-gel, co-precipitation, combustion synthesis, etc. It is expected that in the case of doped Sm2O3, the temperatures required for thermal reduction will be lower as compared to pure Sm2O3. This second phase of this investigation is underway in our laboratories.

4. Conclusions

The thermodynamic feasibility of a Sm-WS cycle for the production of solar hydrogen is investigated. HSC simulation results indicate that the temperature (TH) required for complete thermal reduction of Sm2O3 can be significantly decreased with a reduction in the oxygen partial pressure in the inert flushing gas used inside the reactor. For instance, at oxygen partial pressure in the inert flushing gas = 10−5 bar, TH is 3000 K. As the oxygen partial pressure in the inert flushing gas decreases to 10−8 bar, TH can be lowered to 2280 K. Thermodynamic calculations confirm that the water splitting reaction via Sm oxidation is feasible below 6800 K.
The exergy analysis of the Sm-WS cycle indicates that η a b s o r p t i o n can be increased by 30.61% due to the decrease in TH from 3000 to 2280 K as the oxygen partial pressure in the inert flushing gas reduces from 10−5 to 10−8 bar. In contrast, η a b s o r p t i o n can be reduced by 61.3% with a decrease in the value of C from 10,000 to 2000 suns (at TH = 2280 K). The quantities Q r e a c t o r n e t and Q s o l a r can also be lowered by 4.6 and 39.1% as TH falls from 3000 to 2280 K. Correspondingly, the heat losses due to quenching and re-radiation can be reduced by 120 and 1748 kW for a similar reduction in TH. The lower solar energy input requirements and heat losses during quenching and re-radiation are possible as η a b s o r p t i o n increases with decreasing TH. The maximum efficiencies, η c y c l e = 24.5% and η s o l a r t o f u e l = 29.5%, are observed at lower values of TH = 2280 K and oxygen partial pressure in the inert flushing gas = 10−8 bar (with no heat recuperation). These efficiency values can be further enhanced by recycling the heat recuperated from the quench unit and water splitting reactor to drive the cycle. For instance, at heat recuperation levels of 20%, 40% and 60%, η s o l a r t o f u e l can be increased by a factor of 1.12, 1.28, and 1.49, respectively.

Acknowledgments

This publication was made possible by the NPRP grant (NPRP8-370-2-154) and UREP grant (UREP18-146-2-060) from the Qatar National Research Fund (a member of Qatar Foundation). The statements made herein are solely the responsibility of author(s). The authors also gratefully acknowledge the financial support provided by the Qatar University Internal Grant (QUUG-CENG-CHE-14\15-10).

Author Contributions

This investigation is mainly performed at Qatar University under the guidance of Rahul Bhosale. Other co-authors have contributed in terms of their technical inputs towards the thermodynamic calculations, analysis of the data, and final manuscript preparation.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

CSolar flux concentration ratio (suns)
INormal beam solar insolation (W·m−2)
n ˙ Molar flow rate (mol·s−1)
Q q u e n c h Heat rejection rate to surroundings from quench unit (kW)
Q F C I d e a l Heat rejection rate to surroundings from ideal fuel cell (kW)
Q S m   o x i d a t i o n Heat rejection rate to surroundings from water splitting reactor (kW)
Q S m 2 O 3 h e a t i n g Energy rate required for heating of Sm2O3 (kW)
Q S m 2 O 3 r e d u c t i o n Energy rate required for the thermal reduction of Sm2O3 (kW)
Q r e a c t o r n e t Net energy input rate required for the operation of Sm-WS cycle (kW)
Q r e r a d i a t i o n Radiation heat loss rate from the solar reactor (kW)
Q r e c u p e r a b l e Total heat rate that can be recuperated (kW)
Q s o l a r Solar energy input rate (kW)
Q s o l a r , w i t h   r e c u p e r a t i o n Solar power input after heat recuperation (kW)
THThermal reduction temperature (K)
TLWater splitting temperature (K)
W F C I d e a l Work rate output of ideal fuel cell (kW)
η a b s o r p t i o n Solar absorption efficiency
η c y c l e Cycle efficiency
η s o l a r t o f u e l Solar to fuel energy conversion efficiency
Δ G W S Gibbs free energy change for water splitting reaction (kJ·mol−1)
σ Stefan–Boltzmann constant, 5.670 × 10−8 (W·m−2·K−4)
I r r r e a c t o r Rate of entropy produced across solar reactor (kW·K−1)
I r r q u e n c h Rate of entropy produced across quench unit (kW·K−1)
I r r S m   o x i d a t i o n Rate of entropy produced across water splitting reactor (kW·K−1)

References

  1. International Energy Agency. CO2 Emissions from Fuel Combustion, 2015th ed.; International Energy Agency: Paris, France, 2015. [Google Scholar]
  2. Scheffe, J.R.; Steinfeld, A. Oxygen exchange materials for solar thermochemical splitting of H2O and CO2: A review. Mater. Today 2014, 17, 341–348. [Google Scholar] [CrossRef]
  3. Roeb, M.; Neises, M.; Monnerie, N.; Call, F.; Simon, H.; Sattler, C.; Schmucker, M.; Pitz-Pall, R. Materials-related aspects of thermochemical water and carbon dioxide splitting: A review. Materials 2012, 5, 2015–2054. [Google Scholar] [CrossRef] [Green Version]
  4. Smestad, G.P.; Steinfeld, A. Review: Photochemical and thermochemical production of solar fuels from H2O and CO2 using metal oxide catalysts. Ind. Eng. Chem. Res. 2012, 51, 11828–11849. [Google Scholar] [CrossRef]
  5. Steinfeld, A. Solar hydrogen production via a two-step water-splitting thermochemical cycle based on Zn/ZnO redox reactions. Int. J. Hydrog. Energy 2002, 27, 611–619. [Google Scholar] [CrossRef]
  6. Steinfeld, A. Solar thermochemical production of hydrogen–A review. Sol. Energy 2005, 78, 603–615. [Google Scholar] [CrossRef]
  7. Abanades, S.; Charvin, P.; Flamant, G. Design and simulation of a solar chemical reactor for the thermal reduction of metal oxides: Case study of zinc oxide dissociation. Chem. Eng. Sci. 2007, 62, 6323–6333. [Google Scholar] [CrossRef]
  8. Galvez, M.E.; Loutzenhiser, P.G.; Hischier, I.; Steinfeld, A. CO2 splitting via two-step solar thermochemical cycles with Zn/ZnO and FeO/Fe3O4 redox reactions II: Kinetic analysis. Energy Fuels 2009, 23, 2832–2839. [Google Scholar]
  9. Loutzenhiser, P.G.; Steinfeld, A. Solar syngas production from CO2 and H2O in a two-step thermochemical cycle via Zn/ZnO redox reactions: Thermodynamic cycle analysis. Int. J. Hydrog. Energy 2011, 36, 12141–12147. [Google Scholar] [CrossRef]
  10. Loutzenhiser, P.G.; Meier, A.; Steinfeld, A. Review of the two-step H2O/CO2-splitting solar thermochemical cycle based on Zn/ZnO redox reactions. Materials 2010, 3, 4922–4938. [Google Scholar] [CrossRef]
  11. Dardor, D.; Bhosale, R.R.; Gharbia, S.; Kumar, A.; AlMomani, F. Solar carbon production via thermochemical ZnO/Zn carbon dioxide splitting cycle. J. Emerg. Trends Eng. Appl. Sci. 2015, 6, 129–135. [Google Scholar]
  12. Miller, J.E.; Allendorf, M.D.; Diver, R.B.; Evans, L.R.; Siegel, N.P.; Stuecker, J.N. Metal oxide composites and structures for ultra-high temperature solar thermochemical cycles. J. Mater. Sci. 2008, 43, 4714–4728. [Google Scholar] [CrossRef]
  13. Ishihara, H.; Kaneko, H.; Hasegawa, N.; Tamaura, Y. Two-step water-splitting at 1273–1623K using yttria-stabilized zirconia-iron oxide solid solution via co-precipitation and solid-state reaction. Energy 2008, 33, 1788–1793. [Google Scholar] [CrossRef]
  14. Bhosale, R.R.; Kumar, A.; van den Broeke, L.J.P.; Gharbia, S.; Dardor, D.; Jilani, M.; Folady, J.; Al-Fakih, M.S.; Tarsad, M.A. Solar hydrogen production via thermochemical iron oxide–iron sulfate water splitting cycle. Int. J. Hydrog. Energy 2015, 40, 1639–1650. [Google Scholar] [CrossRef]
  15. Bhosale, R.R.; Kumar, A.; AlMomani, F.A.; Alxneit, I. Sol–gel derived CeO2–Fe2O3 nanoparticles: Synthesis, characterization and solar thermochemical application. Ceram. Int. 2016, in press. [Google Scholar] [CrossRef]
  16. Abanades, S. CO2 and H2O reduction by solar thermochemical looping using SnO2/SnO redox reactions: Thermogravimetric analysis. Int. J. Hydrog. Energy 2012, 37, 8223–8231. [Google Scholar] [CrossRef]
  17. Charvin, P.; Abanades, S.; Lemont, F.; Flamant, G. Experimental study of SnO2/SnO/Sn thermochemical systems for solar production of hydrogen. AIChE J. 2008, 54, 2759–2767. [Google Scholar] [CrossRef]
  18. Dardor, D.; Bhosale, R.R.; Gharbia, S.; AlNouss, A.; Kumar, A.; AlMomani, F. Solar thermochemical conversion of CO2 into C via SnO2/SnO redox cycle: Thermodynamic study. Int. J. Eng. Res. Appl. 2015, 5, 134–140. [Google Scholar]
  19. Bhosale, R.R.; Kumar, A.; Almomani, F. Solar thermochemical hydrogen production via terbium oxide based redox reactions. Int. J. Photoenergy 2016, 2016. [Google Scholar] [CrossRef]
  20. Agrafiotis, C.C.; Pagkoura, C.; Zygogianni, A.; Karagiannakis, G.; Kostoglou, M.; Konstandopoulos, A.G. Hydrogen production via solar-aided water splitting thermochemical cycles: Combustion synthesis and preliminary evaluation of spinel redox-pair materials. Int. J. Hydrog. Energy 2012, 37, 8964–8980. [Google Scholar] [CrossRef]
  21. Bhosale, R.R.; Khadka, R.; Puszynski, J.; Shende, R. H2 generation from two-step thermochemical water-splitting reaction using sol-gel derived SnxFeyOz. J. Renew. Sustain. Energy 2011, 3, 063104. [Google Scholar] [CrossRef]
  22. Scheffe, J.; Li, J.; Weimer, A. A spinel ferrite/hercynite water-splitting redox cycle. Int. J. Hydrog. Energy 2010, 35, 3333–3340. [Google Scholar] [CrossRef]
  23. Bhosale, R.R.; Shende, R.V.; Puszynski, J.A. Thermochemical water-splitting for H2 generation using sol-gel derived Mn-ferrite in a packed bed reactor. Int. J. Hydrog. Energy 2012, 37, 2924–2934. [Google Scholar] [CrossRef]
  24. Bhosale, R.R.; Kumar, A.; AlMomani, F.A.; Alxneit, I. Propylene oxide assisted sol-gel synthesis of zinc ferrite nanoparticles for solar fuel production. Ceram. Int. 2016, 42, 2431–2438. [Google Scholar] [CrossRef]
  25. Roeb, M.; Gathmann, N.; Neises, M.; Sattler, C.; Pitz-Paal, R. Thermodynamic analysis of two-step solar water splitting with mixed iron oxides. Int. J. Energy Res. 2009, 33, 893–902. [Google Scholar] [CrossRef] [Green Version]
  26. Bader, R.; Venstrom, L.J.; Davidson, J.H.; Lipiński, W. Thermodynamic analysis of isothermal redox cycling of ceria for solar fuel production. Energy Fuels 2013, 27, 5533–5544. [Google Scholar] [CrossRef]
  27. Chueh, W.C.; Falter, C.; Abbott, M.; Scipio, D.; Furler, P.; Haile, S.M.; Steinfeld, A. High-flux solar-driven thermochemical dissociation of CO2 and H2O using nonstoichiometric ceria. Science 2010, 330, 1797–1801. [Google Scholar] [CrossRef] [PubMed]
  28. Furler, P.; Scheffe, J.R.; Steinfeld, A. Syngas production by simultaneous splitting of H2O and CO2 via ceria redox reactions in a high-temperature solar reactor. Energy Environ. Sci. 2012, 5, 6098–6103. [Google Scholar] [CrossRef]
  29. Scheffe, J.R.; Steinfeld, A. Thermodynamic analysis of cerium-based oxides for solar thermochemical fuel production. Energy Fuels 2012, 26, 1928–1936. [Google Scholar] [CrossRef]
  30. Scheffe, J.R.; Weibel, D.; Steinfeld, A. Lanthanum–strontium–manganese perovskites as redox materials for solar thermochemical splitting of H2O and CO2. Energy Fuels 2013, 27, 4250–4257. [Google Scholar] [CrossRef]
  31. Demont, A.; Abanades, S. High redox activity of Sr-substituted lanthanum manganite perovskites for two-step thermochemical dissociation of CO2. RSC Adv. 2014, 4, 54885–54891. [Google Scholar] [CrossRef]
  32. Gálvez, M.; Jacot, R.; Scheffe, J.R.; Cooper, T.; Patzke, G.; Steinfeld, A. Physico-chemical changes in Ca, Sr and Al-doped La–Mn–O perovskites upon thermochemical splitting of CO2 via redox cycling. Phys. Chem. Chem. Phys. 2015, 17, 6629–6634. [Google Scholar] [CrossRef] [PubMed]
  33. Roine, A. Outokumpu HSC Chemistry for windows, version 7.1.; Outokumpu Research Oy: Pori, Finland, 2013. [Google Scholar]
  34. Abanades, S.; Charvin, P.; Lemont, F.; Flamant, G. Novel two-step SnO2/SnO water-splitting cycle for solar thermochemical production of hydrogen. Int. J. Hydrog. Energy 2008, 33, 6021–6030. [Google Scholar] [CrossRef]
  35. Diver, R.B.; Miller, J.E.; Allendorf, M.D.; Seigel, N.P.; Hogan, R.E. Solar thermochemical water-splitting ferrite-cycle heat engines. J. Sol. Energy Eng. 2008, 130, 041001. [Google Scholar] [CrossRef]
Figure 1. Typical Sm-WS cycle.
Figure 1. Typical Sm-WS cycle.
Energies 09 00316 g001
Figure 2. Effect of oxygen partial pressure in the inert flushing gas on TH.
Figure 2. Effect of oxygen partial pressure in the inert flushing gas on TH.
Energies 09 00316 g002
Figure 3. Effect of oxygen partial pressure in the inert flushing gas on equilibrium compositions related to the thermal reduction of Sm2O3. Change in the molar concentration of (a): Sm2O3; (b): Sm(g); (c): O2(g).
Figure 3. Effect of oxygen partial pressure in the inert flushing gas on equilibrium compositions related to the thermal reduction of Sm2O3. Change in the molar concentration of (a): Sm2O3; (b): Sm(g); (c): O2(g).
Energies 09 00316 g003
Figure 4. Variation in Δ G w s as a function of TL during H2 production step.
Figure 4. Variation in Δ G w s as a function of TL during H2 production step.
Energies 09 00316 g004
Figure 5. Process configuration for H2 production via Sm-WS cycle.
Figure 5. Process configuration for H2 production via Sm-WS cycle.
Energies 09 00316 g005
Figure 6. Effect of (a) TH at C = 10,000 suns; and (b) C on η a b s o r p t i o n at TH = 2280 K.
Figure 6. Effect of (a) TH at C = 10,000 suns; and (b) C on η a b s o r p t i o n at TH = 2280 K.
Energies 09 00316 g006
Figure 7. Effect of TH on Q s o l a r and Q r e a c t o r n e t .
Figure 7. Effect of TH on Q s o l a r and Q r e a c t o r n e t .
Energies 09 00316 g007
Figure 8. Effect of TH on I r r r e a c t o r and I r r q u e n c h .
Figure 8. Effect of TH on I r r r e a c t o r and I r r q u e n c h .
Energies 09 00316 g008
Figure 9. Effect of % heat recuperation on Q s o l a r ,   w i t h   r e c u p e r a t i o n and Q r e c u p e r a b l e (TH = 2280 K).
Figure 9. Effect of % heat recuperation on Q s o l a r ,   w i t h   r e c u p e r a t i o n and Q r e c u p e r a b l e (TH = 2280 K).
Energies 09 00316 g009
Figure 10. Effect of % heat recuperation on η c y c l e and η s o l a r t o f u e l (TH = 2280 K).
Figure 10. Effect of % heat recuperation on η c y c l e and η s o l a r t o f u e l (TH = 2280 K).
Energies 09 00316 g010
Table 1. Variation in Q q u e n c h and Q r e r a d i a t i o n as a function of TH.
Table 1. Variation in Q q u e n c h and Q r e r a d i a t i o n as a function of TH.
TH (K)Qquench (kW)Qre−radiation (kW)
30007562194
27807141301
2540673773
2280636446
Table 2. Values for η c y c l e and η s o l a r t o f u e l as a function of TH.
Table 2. Values for η c y c l e and η s o l a r t o f u e l as a function of TH.
TH (K)ηcycle (%)ηsolar−to−fuel (%)
300014.917.9
278018.522.3
254021.726.2
228024.529.5
Table 3. η c y c l e and η s o l a r t o f u e l for Sm-WS cycle.
Table 3. η c y c l e and η s o l a r t o f u e l for Sm-WS cycle.
TH (K)ηcycle (%)ηsolar−to−fuel (%)
Recuperation = 0%
278018.522.3
254021.726.2
228024.429.5
Recuperation = 20%
278020.324.5
254024.129.1
228027.533.1
Recuperation = 40%
278022.427.1
254027.232.8
228031.437.8
Recuperation = 60%
278025.130.3
254031.137.5
228036.544.1
Table 4. Comparison between Sm-WS cycle and previously investigated metal oxide based cycles in terms of η s o l a r t o f u e l .
Table 4. Comparison between Sm-WS cycle and previously investigated metal oxide based cycles in terms of η s o l a r t o f u e l .
Cycleηsolar−to−fuel (%)
Zinc oxide cycle29.0
Tin oxide cycle29.8
Iron oxide cycle30.0
Ceria cycle20.2
Sm-WS cycle44.1

Share and Cite

MDPI and ACS Style

Bhosale, R.; Kumar, A.; AlMomani, F.; Ghosh, U.; Saad Anis, M.; Kakosimos, K.; Shende, R.; Rosen, M.A. Solar Hydrogen Production via a Samarium Oxide-Based Thermochemical Water Splitting Cycle. Energies 2016, 9, 316. https://doi.org/10.3390/en9050316

AMA Style

Bhosale R, Kumar A, AlMomani F, Ghosh U, Saad Anis M, Kakosimos K, Shende R, Rosen MA. Solar Hydrogen Production via a Samarium Oxide-Based Thermochemical Water Splitting Cycle. Energies. 2016; 9(5):316. https://doi.org/10.3390/en9050316

Chicago/Turabian Style

Bhosale, Rahul, Anand Kumar, Fares AlMomani, Ujjal Ghosh, Mohammad Saad Anis, Konstantinos Kakosimos, Rajesh Shende, and Marc A. Rosen. 2016. "Solar Hydrogen Production via a Samarium Oxide-Based Thermochemical Water Splitting Cycle" Energies 9, no. 5: 316. https://doi.org/10.3390/en9050316

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop