Next Article in Journal
Research on Green Design Strategy of Electrical and Electronic Manufacturing Enterprises Based on the Perspective of Tripartite Evolutionary Game
Previous Article in Journal
Agro-Food Supply Chains in Peri-Urban Agricultural Areas: Do They Contribute to Preserve Local Biodiversity? The Case of Baix Llobregat Agrarian Park
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Research in Solar-Driven Hydrogen Production

by
Yimin Deng
1,
Shuo Li
2,
Helei Liu
1,
Huili Zhang
2 and
Jan Baeyens
3,*
1
School of Chemistry and Chemical Engineering, Beijing Institute of Technology, Beijing 102488, China
2
School of Life Science and Technology, Beijing University of Chemical Technology, Beijing 100029, China
3
Process Technology and Environmental Lab, Department of Chemical Engineering, KU Leuven, 2860 Sint-Katelijne-Waver, Belgium
*
Author to whom correspondence should be addressed.
Sustainability 2024, 16(7), 2883; https://doi.org/10.3390/su16072883
Submission received: 21 February 2024 / Revised: 25 March 2024 / Accepted: 27 March 2024 / Published: 29 March 2024

Abstract

:
Climate concerns require immediate actions to reduce the global average temperature increase. Renewable electricity and renewable energy-based fuels and chemicals are crucial for progressive de-fossilization. Hydrogen will be part of the solution. The main issues to be considered are the growing market for H2 and the “green” feedstock and energy that should be used to produce H2. The electrolysis of water using surplus renewable energy is considered an important development. Alternative H2 production routes should be using “green” feedstock to replace fossil fuels. We firstly investigated these alternative routes through using bio-based methanol or ethanol or ammonia from digesting agro-industrial or domestic waste. The catalytic conversion of CH4 to C and H2 was examined as a possible option for decarbonizing the natural gas grid. Secondly, water splitting by reversible redox reactions was examined, but using a renewable energy supply was deemed necessary. The application of renewable heat or power was therefore investigated, with a special focus on using concentrated solar tower (CST) technology. We finally assessed valorization data to provide a tentative view of the scale-up potential and economic aspects of the systems and determine the needs for future research and developments.

1. Introduction

1.1. Reviewing the Current H2 Production Routes

The current energy systems are environmentally unsustainable with reduced resources, increasing the world’s population and increasing energy use, hampering the energy future. Future energy supplies will need to apply clean and CO2-lean renewable sources whilst partly decarbonizing the energy sector. Regarding the target objectives, the major energy consumers will be the building, industry, and transportation sectors. The intermittent nature of some renewables (wind, solar) will result in the use of energy storage techniques. Novel techniques will need to be developed to manufacture environmentally friendly energy carriers that can be easily stored and transported [1].
Producing H2 has a high potential to either decarbonize the natural gas grid [2], to directly drive fuel cells, to reduce iron ore [3], or to fire cement, ceramic, or limestone kilns [4,5]. The current methods used for the production of H2, namely the reforming of fossil feedstock or H2 production by “renewable” routes of biomass gasification and carbohydrate fermentation, still release CO2. These routes have been previously assessed [6]. Novel approaches will deal with “green” feedstock such as CH4 (digester biogas), NH3 (stripping of digestate), methanol, and ethanol (fermentations), and even their fossil fuel-based equivalents can be considered. These feedstock materials can mostly be catalytically steam reformed into H2. The decomposition of CH4 and NH3 does not require H2O addition. Thermal water decomposition by oxidation reduction reactions has been investigated but is not fully proven yet [6].
To reduce CO2 emissions, Abdin et al. estimate that approximately 1 billion tons of non-fossil-based H2 is required, 10 times more than the current amount of H2 produced [7]. Although electrolysis using renewable energy is promoted, the current supplies are insufficient to meet this goal [8].
H2 is currently mainly (~50%) produced by the steam reforming of natural gas/oil/naphtha. About 30% is covered by the refinery/chemical industry, and coal gasification produces about 18%. Water electrolysis (3.9%) and other minor sources make up the rest of the annual H2 production. In the future, “green” H2 processes will need to supply H2 for traditional and new applications [9,10].
The current H2 applications are mainly in oil refineries and in chemicals’ syntheses, from the hydrogenation of fats/oils to the manufacturing of pharmaceuticals, among others. The future is expected to enhance the demand for H2 for both chemical processes and for transportation fuel.
The different current ways to produce H2 have limitations and advantages, as summarized in Table 1 [11,12]. The tentative cost of H2 varies among the processes, which have different levels of maturity. The specific case of electrolysis is illustrated in Table 2.
Regarding electrolysis, Table 2 refers to mature alkaline electrolysis [13]. The proton exchange [H+] polymer electrolyte membrane (H+-PEM) is ready for commercial application for operation at 20 to 200 °C, with an estimated efficiency of 65 to 82% [14,15,16]. Solid oxide electrolysis (SOE) processes can reach a 100% efficiency while operating at high temperatures [17]. While the H+-SOE type is close to the demonstration scale, the O2−-SOE is still at a laboratory scale. The H2 production costs associated with H+-PEM and SOE have not yet been fully determined, but cheap photovoltaic or curtailed wind energy (<0.02 USD/kWh) will be required to achieve costs between 1.4 and 1.7 USD/kg H2. The required 0.02 USD/kWh is far below the current cost.
Table 2. Characteristics of electrolysis systems for H2 generation.
Table 2. Characteristics of electrolysis systems for H2 generation.
TechnologyAlkaline [13]PEM [14]SOE [18]
Technology MaturityMatureDemonstrationR&D
Temperature, °C60–8050–80500–1000
Pressure, bar<30<30<30
Energy consumption, kWh/Nm34.5–7.04.5–7.52.5–3.5
Surface area of the cells, m2<4<300-
H2 yield, Nm3/h<760<30-
Tentative lifetime, yr20–3010–20-
Hydrogen quality, %>99.8>99.99-

1.2. The Current Development of Electrolysis

Although only responsible for ~4% of the current amount of H2 produced, electrolysis developments are important. Whatever developments are selected, the process will require cheap electricity and should hence be considered in conjunction with nuclear power, curtailed PV or wind power, or concentrated solar power. The construction of the process electrodes is a major research topic. De-ionized water is required before electrolysis, and catalysts (e.g., platinum) will be required. The reaction is straightforward and will decompose water (liquid) into H2 (g) and O2 (g) at an E0 value of −1.229 V.
The current cost of H2 production through electrolysis largely exceeds the cost of the reforming production methods, and future progress is only guaranteed if the electrolysis process applies cheap renewable electricity. Table 2 summarizes water electrolysis developments.
The immediate future of green electrolytic H2 relies mostly on wind or photovoltaic electricity despite the high specific energy required [8]. The further growth of electrolysis will not be capable of meeting the expected 1-billion-ton H2 goal. Fossil-based feedstock will hence remain the dominant H2 source despite the CO2 emissions of about 10 kg of CO2/kg H2 for steam methane reforming (“gray hydrogen”) [19]. The technology could be “cleaner” if carbon capture and storage were applied. This would, however, involve significant costs, nearly double the produced H2 costs [20].

2. Current Hot-Topic H2 Production Methods

2.1. Thermal Single-Step Water Splitting

Water can be thermally dissociated (water thermolysis) in H2 (g) and 0.5 O2 (g). To achieve a fair water dissociation, high temperatures above 2500 K are required [21]. Moreover, the process requires the efficient separation of H2 and O2 to avoid explosion hazards. Membranes with, e.g., ZrO2 can be used up to 2500 K. Separation can also be achieved at low temperatures after quickly (millisecond level) quenching the product gas mixture to below 500 K, where Palladium membranes can be efficiently used. Only nuclear or concentrated solar towers can meet the required high temperatures, but the selection of usable high-temperature construction materials remains a major bottleneck [22].

2.2. Thermo-Chemical Redox Water Splitting

Most of the thermo-chemical redox cycles published to date belong to one of five generic cycles shown in Table 3 and Figure 1 [23,24]. M is either in pure metal form or its lower valency cation in an oxide or halide pair.
The simpler schemes are desirable in terms of efficiency and capital cost. Only a few redox cycles can operate below a temperature of 1000 K. Many compounds form multi-step cycles and require a high operating temperature. Moreover, the incomplete reversibility reduces the process efficiency [23,24]. Possible developments are further discussed in subsequent sections.

2.3. Conversion of Methane (or Hydrocarbons)

Commercial bulk H2 is commonly produced from methane or hydrocarbons. These processes have a long history (>50 years). Steam methane reforming (SMR) operates at 700 to 1100 °C over a metal-based catalyst (nickel). Steam reacts with methane to produce CO and H2.
C H 4 + H 2 O C O + 3 H 2
C O + H 2 O C O 2 + H 2
Both reactions are reversible [25], and the process efficiency is between 65 and 75%. The CO2 must be abated. The alternative, the direct conversion of CH4, requires temperatures above 1500 K. The occurring reactions are given below.

2.4. Methanol to Hydrogen

Due to the difficulties in transporting and handling H2, methanol, containing 12.6 wt% of H, is considered as a cheap and abundant feedstock. Although methanol is mostly produced by chemical synthesis, new “green” production methods have emerged through biochemical methods.

2.5. NH3 to Hydrogen

The reversible decomposition reaction of ammonia follows the below-listed equation:
NH 3 0.5   N 2 + 1.5   H 2   H 0 = 46.2   k J / m o l
The decomposition reaction is mildly endothermic and needs a catalyst to lower the activation energy. Many possible catalysts can be used for this reaction. Several catalysts were investigated in the present study, and we will report the function of different temperatures and reaction times. Details are reported in Section 4.

2.6. Objectives of the Research

Further reported research will also demonstrate that fairly low-temperature catalytic steam reforming is possible, as initially proposed in [25,26].
The results of our research demonstrate that decentralized H2 production from transportable and cheap CH3OH is possible in a low T, P process using durable and selective nanoparticle-based catalysis, with low-to-no CO being produced. This method has potential applicability in the transportation of fuel cells. Finally, redox-driven water splitting will be illustrated, together with some recent developments in using concentrated solar heat to meet endothermic reaction requirements.

3. Solar H2 Production

3.1. Near-Term Options

Since all conversion processes are endothermic, a cheap energy source is required, with wind and solar energy being used in the energy supply of endothermic reactions. Figure 2 and Table 4 illustrate the different routes for producing H2 by thermolysis, electrolysis, and photonic dissociation. Concentrated solar power (CSP) systems can be applied in the solar production of H2. Figure 2 illustrates how concentrated solar radiation can be used in the solar production of H2 from water or carbonaceous materials.
Water, as an abundant and cheap reactant, should be the ultimate H2 feedstock. It is carbon-free and can be decomposed through solar-driven water splitting by thermo-chemical redox cycles. Its large-scale application will depend upon the achievement of a fairly low temperature and multi-cycle reversibility [18,27]. Since solar-driven systems still suffer from a low cost-effectiveness and intermittency, novel solutions are required to increase their efficiency and provide a considerable cost reduction [28]. Even in solar–thermal applications, the catalytic steam reforming of low-cost biochemicals such as bio-ethanol has a high potential, with a H2 production efficiency exceeding 95%.
Table 4 summarizes the characteristics of different reported solar H2 processes. Although full data have not yet been given, a positive appraisal emerges from the comparison. These processes can foster the future production of “green” hydrogen.
Moreover, it should not be forgotten that hybrid solar PV and on/offshore wind systems are currently being investigated, demonstrating that the cost in USD/kg H2 will exceed the current petrochemical production costs, except in parts of the world with a high PV and wind electricity potential. This is illustrated in Figure 3.
Table 4. Characteristics of developed solar H2 production techniques.
Table 4. Characteristics of developed solar H2 production techniques.
Solar SystemReqd. TTypeFeedCO2 ProductSystem Efficiency (%)Capacity, (ton H2/d)H2 Price, (USD/kg)
Concentrated solar thermal
Solar tower [29,30,31]LowElectrolysis (PEM)WaterNo2138.355.1
HighThermolysisCoalCCS-integrated38–40--
Solar dish [32]HighSteam electrolysisWaterNo201.3610.5
Parabolic trough [32]HighSteam electrolysisWaterNo2062.956.5
Concentrated solar tower [33,34,35]HighThermochemical cycleHydrocarbonsNo-68.0
HighThermochemical cycleWater/CO2No4.1--
HighThermochemical cycleWater/CO2No22.8–39.6--
Concentrated solar tower [4,36]HighThermolysisWaterNo29.93320 L/h
HighGasificationSolid carbonaceous (e.g., coal)Yes---
HighPyrolysisNG, oil, and hydrocarbonsYes---
HighSteam reformingNG, oil, and hydrocarbonsYes---
HighThermolysisWaterNo5.6--
PV [28]LowElectrolysis (PEM)WaterNo9.81012.1
Photo-electrochemical [37]HighPhoto-catalysisWaterNo---
Solar power plant [38]HighSolar powerWaterNo24.9--
Hybrid systems [31,38,39,40,41]
Solar thermal [42]HighElectrolysis (PEM)WaterNo10716.3
PV–grid [28]LowElectrolysis (PEM)WaterNo9.8106.1
PV–thermal [39,41]HighThermolysisWaterNo---
HighThermolysisVariousLimited---
PV–CSP [30]HighThermolysisWaterNo--2.7
Solar and wind [40]HighElectrolysisVariousLimited---

3.2. The Concentrated Solar Tower Solution

Concentrated solar energy is of increasing interest in power and industrial heat generation. The intermittency of solar energy necessitates the integration of a heat storage system to operate on a continuous basis. The applications of different CSP technologies are mainly a function of the targeted temperature ranges and the intended applications, as illustrated in Figure 4.
For solar power applications, both parabolic trough and central tower concepts are mostly used, with parabolic troughs being responsible for most of the moderate temperature thermodynamic cycles. Central tower concepts, however, have gained importance as Peaker solutions since new heat transfer media (i.e., particle suspensions) enable a higher temperature application, hence fostering the use of high-efficiency thermodynamic cycles.
Such a particle-driven concept has been developed over the past few years with a particle loop operation at a temperature around 800 °C. A vertical multi-tube (40) solar receiver reached a capture yield of >2 MWth. The future scale-up to between 10 and 50 MWth will apply the same multi-tube layout with taller tubes of ≥6 m.

4. Experimental Investigation of H2 Production from “Gray” or “Green” Feedstock

Four experimental set-ups were used, each of which are depicted in Figure 5. The reactors contained fixed or vibro-fluidized beds of the appropriate catalyst particles (15 cm bed depth for the fixed bed mode in the electrical furnaces and 25 cm in the sun-heated fluidized bed reactor). The general layout of the experiments is given in Figure 5a, while equipment details are given in Figure 5b–e. Details of the rigs are presented in Deng et al. [2,44].
The first “gray” feedstock, CH4, was investigated mostly for its hydrogen-enhanced natural gas (HENG) potential. Our experimental investigation applied wet-impregnated Fe/α-Al2O3 and Ni/MgO catalysts. The 20 wt% Fe/α-Al2O3 catalyst yields nearly complete CH4 conversion at 700 °C and is preferred for CH4 conversion. Ni/MgO catalysis leads to lower conversions at similar operating conditions (Figure 6). The decomposition carbon accumulates on and within the catalyst, progressively reducing the H2 yield. A periodic regeneration (burning off the coke deposit) is required after 20 to 30 h of production. At temperatures below 635 °C, the reaction kinetics are dominated by the reaction rate. Mass transfer limitations are increasingly important at higher temperatures. Figure 6 depicts the conversion effects of reaction time and temperature. The use of the Fe catalyst is recommended.
A second “gray” feedstock (with the potential to become “green” through future developments in the bio-synthesis of methanol feedstock), CH3OH, was examined. The experimental results in Figure 7 show that the yield of the catalytic steam reforming of methanol is almost 2.5 mol H2/mol CH3OH (slightly lower than the stoichiometric yield of 3 mol H2/mol CH3OH). The CH3OH–carbon is mostly detected as CO and CO2, as illustrated in Table 5. No coke deposits were detected after long-duration experiments. The catalysts maintained their selectivity and activity. The kinetics demonstrated a high reaction rate. The MnFe2O4 catalyst is four times slower and also requires a higher conversion temperature (400 °C).
Using the Co/α-Al2O3 catalyst is recommended for scale-up and fuel cell application due to the higher H2 yield and lower CO formation.
Thirdly, the production of H2 from “green” or industrial NH3 was studied in the present research using non-noble metal catalysts. “Green” ammonia sources were investigated, such as ammonia produced by stripping from manure and digestate. Three catalysts were tested: wet-impregnated Fe-Al2O3, dry-milled Fe-Al2O3, and wet-impregnated Ni-Al2O3. The results are illustrated in Figure 8. The mildly endothermic decomposition of NH3 into H2 and N2 is achieved with a 100% yield at 500 °C using a Fe/α-Al2O3 catalyst. The alternative and more expensive Ni/α-Al2O3 catalyst scores significantly lower. The experimental results were converted into reaction kinetics. The reaction is very fast, with a rate constant of 7.5/s at 500 °C. Full conversion is obtained within 0.4 s. Very pure (>95%) H2 can be produced and stored using several techniques. It is proven that the production of H2 from NH3 has a high potential and merits pilot-scale consideration.
For the Catalytic steam reforming of ethanol (CSRE), the most promising catalysts were selected from an in-depth literature review and tested in the present research using all the equipment alternatives in Figure 9, and this test was carried out for more than 30 successive hours. The H2 production efficiencies (5.5 mol H2/mol ethanol) were close to the expected stoichiometric values. CO and CO2 were the main carbon-bearing reaction products. The results are illustrated in Figure 9. A kinetic analysis defined the reaction rate expressions, and the reaction rate per unit catalyst weight was established at 0.58 mol H2/gcat h. Abundant bio-ethanol can be obtained from first- and second-generation fermentation.
The average yields, expressed as mol products/mol EtOH, for 30 h of continuous operation were 5.463 ± 0.306 for H2, 0.670 ± 0.103 for CO, 0.038 ± 0.012 for CH4, and 1.196 ± 0.063 for CO2 in the operation mode of Figure 9a. For the operation mode of Figure 9b, the results were 5.149 ± 0.276 for H2, 0.579 ± 0.073 for CO, 0.047 ± 0.024 for CH4, and 1.123 ± 0.100 for CO2.

5. Water Splitting by Redox Reactants

While water splitting by thermolysis occurs at very high temperatures, redox-driven water splitting cycles can operate at moderate temperatures. Based on previous studies in the literature, 24 investigated cycles, mostly investigated by small-scale experiments, were examined by a multi-criteria decision-making exercise [18]. From the objective screening, the priority redox systems selected used MnOx/Na2CO3 and MnFe2O4/Na2CO3. These systems were those further investigated for short- and long-duration cycling [18].
The thermochemical cycle of MnFe2O4/Na2CO3 is generally described by the following reactions:
2 M n F e 2 O 4 s + 3 N a 2 C O 3 s + H 2 O ( g ) 6 N a ( M n 1 / 3 F e 2 / 3 ) O 2 ( s ) + 3 C O 2 ( g ) + H 2 ( g )
6 N a ( M n 1 / 3 F e 2 / 3 ) O 2 s + 3 C O 2 g 2 M n F e 2 O 4 s + 3 N a 2 C O 3 s + 0.5 O 2 ( g )
The thermochemical cycle of MnOx/Na2CO3 is generally described by the following reactions:
3 N a 2 C O 3 s + 2 M n 3 O 4 s 4 N a M n O 2 s + 2 C O 2 g + 2 M n O s + N a 2 C O 3 s
2 M n O s + N a 2 C O 3 s + H 2 O v H 2 g + C O 2 g + 2 N a M n O 2 s
6 N a M n O 2 s + a y H 2 O v + ( 3 + b ) C O 2 g 3 N a 2 C O 3 s + a H x M n O 2 · y H 2 O s + b M n C O 3 s + c M n 3 O 4 s
a H x M n O 2 · y H 2 O s + b M n C O 3 s 2 c M n 3 O 4 s + a y H 2 O v + b C O 2 g + 0.5 O 2 g
For the MnOx/Na2CO3 redox water splitting system, tests were carried out in the electrical furnace (Figure 5b) and in the solar furnace (Figure 5d) at 775 and 825 °C, with 10 and 250 g of redox reactant, respectively. Solar results are illustrated in Figure 10. Milled particles (<1 µm) with a large surface area provide better results. Ten complete oxidation/reduction cycles were tested. The cumulative H2 production achieved 78% at 825 °C. Milled Mn3O4/Na2CO3 reached a yield of over 80% in the first cycle and over 77% after 10 cycles. The Mn3O4/Na2CO3 system obtained an efficiency of >95% at 775 °C in the solar furnace despite the high T required, but provided that cheap thermal energy could be supplied, over 100 cycles are needed to reach H2 production costs below 2 EUR/kg. The separation and reused of CO2 in the multi-step reaction system remain a major drawback.
The MnFe2O4/Na2CO3 cycles were isothermally tested for more than 200 successive cycles by thermogravimetry in the electric furnace and in the solar reactor. A H2 production efficiency in excess of 85% was obtained with the co-precipitated reactant after multiple cycles. The results are presented in Figure 11. Our solar tests confirmed >95% H2 efficiency. The reverse CO2 reaction time should exceed 4.5 h; 6 h is preferable. The CO2 for the regeneration step was introduced after each oxidation (H2 release) step. CO2 was not recovered in the experiments, and bottled CO2 was used. The preliminary results with a reduction step of 3 h [45] demonstrated that the regeneration was incomplete. Only a regeneration duration of 4.5 (preferably 6) hours could nearly regenerate the redox reactant and achieve a high H2 yield after multi-cycle reactions. Tentative economic calculations revealed that 120 consecutive cycles would produce H2 at 1 EUR/kg H2. However, this is only possible if a cheap energy supply is available and if CO2 can be fully reused in the reduction reaction. The 30% cheaper ball-milled reactant will proportionally reduce costs. The obtained results demonstrate that the further improvement of the system is important.
The H2 production efficiency can largely exceed 90% when appropriate oxidation and reduction times are applied. Co-precipitated MnFe2O4 costs ~500 EUR/ton, as opposed to ~350 USD/ton for the ball-milled alternative.
Both the use of renewable energy and good heat management amongst the cycles of the system are important for limiting heating costs. Since CO2 is required in the reverse reaction, it should be fully recovered. The proposed MnFe2O4/Na2CO3 cycle needs further testing at pilot scale to really prove its competitiveness with other H2 production methods with the electrolysis of water using solar-based electricity or heat. Pilot- and commercial-scale experiments are required to confirm lab-scale results.

6. Valorization Potential

In this section, special attention is paid to the scale-up and Technology Readiness Level of the investigated H2 production technologies, their costs, and associated process requirements, such as H2 upgrading, the potential solid oxide fuel cell applications, and additional technical matters. Essential findings are summarized and used in scale-up procedures for both the thermo-catalytic conversion of CH4, CH3OH, C2H5OH, and NH3 and for MnFe2O4/Na2CO3 redox water splitting. The additional topics of H2 upgrading, the safety issues associated with H2 production and its use, and high-temperature gas filtration are discussed. A techno-economic assessment and a tentative discussion of the Technology Readiness Level (TRL) will be presented. Finally, the solar receivers will be briefly discussed according to their temperature classification.
Thermo-catalytic H2 production: The experimental results provided design equations and methods for these reaction systems. However, a distinction needs to be made between the CH4 feedstock and the CH3OH, NH3, and C2H5OH feedstock. In the case of CH4, coking is a major issue, with the catalyst being progressively de-activated if not thermally regenerated. The reaction temperatures are also moderate to high. The catalyst is an A-type particle, meaning it is easily fluidized and conveyed. No H2O is required. In the other cases, no coking is observed, and low (250 °C) to moderate (500 °C) temperatures are required. The catalyst is, however, of μm grade. This results in very high reaction rates, near-complete conversions at short contact times, and high gas hourly space velocities (GHSVs), as listed in Table 6.
Small amounts of catalyst can generate significant amounts of H2. These differences will significantly affect the scale-up potential, as illustrated below. In view of the catalyst size, a fluidized bed (CH4, NH3) or a fixed bed (methanol, ethanol) were proposed and studied using different plant layouts. The main layout is depicted in Figure 12.
H2 upgrading: As experimentally demonstrated, a C-H-O feedstock will always lead to the formation of reaction byproducts, mostly CO and CO2. The produced gas will need to be treated in a post-treatment for the separation and recovery of individual compounds. Figure 13 depicts a separation unit successively removing H2O vapor, removing CO2, and carrying out the membrane recovery of H2. Current membrane modules have a proven H2 purity and recovery >95%. The detailed design of the upgrading was presented by Deng et al. [44].
At such high recovery, minor amounts of H2 will accompany the CO and CH4 flows, and these flows will be the potential heat source for the feedstock evaporation and preheating.
Safety issues of producing/using H2: The literature on the safety issues associated with producing and using hydrogen is extensive, and attention is specifically paid to the selection of high-pressure durable materials to counter hydrogen attack, embrittlement, and cracking. The proposed materials differ from high to low pressure storage. While austenitic stainless steel, Al alloys, Cu, and Cu alloys are recommended for their high tensile strength at high pressures, low density, and chemical H2 inertness, low-pressure vessels can be constructed in AISI 316, 304, 304 L, and 316 L steels. Cu alloys and Al alloys can also withstand H2 attack.
Redox-driven water splitting results demonstrate that MnFe2O4/Na2CO3 is the ideal redox pair, with a recommended 3 h oxidation cycle (with H2O) and a 6 h reduction cycle (with collected CO2). Efficiencies, even after 35 cycles, reach above 90% at 600 °C. Although experiments with MnFe2O4 were only conducted for long durations at TGA scale, preliminary solar-driven tests were performed during the oxidation cycle. The maximum temperature of the reactor wall in the cavity must be limited to 900 °C (thermo-mechanical strength limits of Incoloy); the bed temperature was kept uniform between 710 and 735 °C by heliostat focusing. The solar H2 yield at 735 °C was nearly the theoretical H2 yield of 0.5 mol/mol, as shown in Figure 14. Since the operation requires a twin-reaction mode (oxidation/reduction), both reactors were installed inside the CST cavity. The reactors will be operated in a reversible mode during consecutive cycles (and probably consecutive days). Since the reactor will be of a fixed-bed nature, either a concentrated solar tower or a parabolic dish solar concept can be used.
The Technology Readiness Level (TRL) reflects the development level of a given technology based upon various parameters determined for long-duration applications, such as robustness, reliability, production costs, energy efficiency, and utilization constraints, among others. A hybrid operation mode, with curtailed wind/PV/concentrated solar supplies, seems the obvious solution. Together with heat management within the process, concentrated solar energy from the receiver should be stored and used in periods where cheaper PV electricity is not available (in non-sun periods). Whereas catalytic conversions of appropriate feedstock operate at moderate temperatures, the significantly higher temperatures needed in water splitting are more problematic. At the current stage of thermo-catalytic or thermo-redox H2 production, it must be admitted that they only achieve a TRL between 1 and 3.
Future cost estimations are extremely challenging, subjective, and inaccurate, mostly because of the low TRLs of the processes. Cost estimations are hence premature. While no solar H2 production method has achieved a fair TRL so far, further research, preferably pilot- or large-scale research, is urgently needed to help establish the ongoing hydrogen revolution. At present, the state-of-the-art systems only achieve a TRL between 1 and 3.
Techno-economic assessment: Table 1 already included a cost comparison of the different H2 production methods. A further comparison was made on the basis of the levelized cost of hydrogen (LCOH) and the calculated GHG emissions. Table 7 illustrates this comparison. Developed catalytic feedstock decomposition and redox water splitting offer major advantages.
Redox water splitting using the MnFe2O4/Na2CO3 system could achieve a low LCOH (<1 USD/kg H2) for recyclability for over 120 cycles.
Solar energy as a heat source in particle-driven systems: The Next-CSP receiver is of specific interest in the concentrated solar power technology. Its potential capacity up to 50 MWth, however, exceeds the needs of H2 production concepts where 100 kWth to 1 MWth would meet the required needs. It is hence important to stress the applicability of two cheaper concentration technologies: a crossflow fluidized bed receiver, which is installed in the focusing cavity, and the improved solar dish technology, developed by the Chinese Academy of Sciences. Both are currently being tested for water splitting and catalytic ethanol conversions. The small-footprint solar dish system can reach very high temperatures (dictated by the construction materials) and operates at a solar concentration ratio of around 2000.
A 64 m2 solar dish is depicted in Figure 15. Ongoing developments will help reach 500 m2, thus allowing for the capturing of 500 kWth at a DNI of 1 kW/m2. With a 790% dish-to-cavity efficiency and a 75% receiver efficiency, up to ~260 kWth can be used in the ethanol or water splitting reactions.
The 2025 UN sustainable development goals (SDGs) comprise 17 objectives to be achieved by 2030. The above assessment of the “green” H2 production methods demonstrates that solar or curtailed renewable energy processes will significantly contribute to meeting these goals [44].

7. Conclusions and Recommendations

The general conclusions of this research review are summarized below:
(i)
Renewable energy-based H2 production from “green” and “gray” feedstock in the presence of steam: the reaction is fast; first-order kinetics and operation at T << 600 °C; near-stoichiometric H2 yields are achieved; CO, CO2, and CH4 are closing the C balance.
(ii)
Two priority water splitting redox reactions are favored: MnOx/Na2CO3 is possible at high T, but recyclability is incomplete. MnFe2O4/Na2CO3 has a high potential. Recyclability for >35 cycles was proven. If >120 cycles can be achieved, the H2 cost will be <1 USD/kg H2.
(iii)
Thermo-catalytic and thermo-chemical processes can work with heating by excess wind/PV power, preferably with CST for continuous operation.
(iv)
Bio-ethanol is attractive since we can accept crude bio-ethanol (after filtration), thus avoiding two distillation steps and the final molecular sieve dewatering, so the price reduction is about 30%.
(v)
Ammonia must be further developed since it may be a solution for on-car H2 generation (ammonia is easy to store). This could also offer possibilities for methanol.
(vi)
The design recommendations and tentative economic balances should be proven at pilot scale. This experimental research study provides conclusions that are positive enough to inspire other researchers and even R&D centers to continue this research.

Author Contributions

Methodology, Y.D.; validation, S.L. and H.Z.; data curation, Y.D. and S.L.; writing—original draft preparation, Y.D., J.B. and H.Z.; writing—review and editing, Y.D., J.B., H.L. and H.Z.; supervision, J.B.; project administration, H.L.; funding acquisition, H.L. All authors have read and agreed to the published version of the manuscript.

Funding

The authors gratefully acknowledge the support of the National Natural Science Foundation of China (NSFC-No. 22208017).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The raw data supporting the conclusions of this article will be made available by the authors on request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pekic, S. Facts and Factors: Green Hydrogen Market to Grow to $1.423m by 2026; Offshore Energy: Schiedam, The Netherlands, 2021. [Google Scholar]
  2. Deng, Y.; Dewil, R.; Appels, L.; Van Tulden, F.; Li, S.; Yang, M.; Baeyens, J. Hydrogen-Enriched Natural Gas in a Decarbonization Perspective. Fuel 2022, 318, 123680. [Google Scholar] [CrossRef]
  3. Li, S.; Zhang, H.; Nie, J.; Dewil, R.; Baeyens, J.; Deng, Y. The Direct Reduction of Iron Ore with Hydrogen. Sustainability 2021, 13, 8866. [Google Scholar] [CrossRef]
  4. CN CemNet. Buxton Plant Manufactures Industrial Lime with Hydrogen Technology. Available online: https://www.cemnet.com/News/story/173001/buxton-plant-manufactures-industrial-lime-with-hydrogen-technology.html (accessed on 22 July 2022).
  5. Lansdorf, T. Hydrogen—A Game Changer for the Ceramic Industry. Interceram-Int. Ceram. Rev. 2022, 71, 48–54. [Google Scholar] [CrossRef]
  6. Chelvam, K.; Hanafiah, M.M.; Woon, K.S.; Ali, K. Al A Review on the Environmental Performance of Various Hydrogen Production Technologies: An Approach towards Hydrogen Economy. Energy Rep. 2024, 11, 369–383. [Google Scholar] [CrossRef]
  7. Antweiler, W. What Role Does Hydrogen Have in the Future of Electric Mobility? Available online: https://wernerantweiler.ca/blog.php?item=2020-09-28 (accessed on 4 January 2024).
  8. Abdin, Z.; Zafaranloo, A.; Rafiee, A.; Mérida, W.; Lipiński, W.; Khalilpour, K.R. Hydrogen as an Energy Vector. Renew. Sustain. Energy Rev. 2020, 120, 109620. [Google Scholar] [CrossRef]
  9. BP p.l.c. Hydroelectricity. Available online: https://www.bp.com/en/global/corporate/energy-economics/statistical-review-of-world-energy/hydroelectricity.html (accessed on 4 January 2024).
  10. Minutillo, M.; Perna, A.; Sorce, A. Combined Hydrogen, Heat and Electricity Generation via Biogas Reforming: Energy and Economic Assessments. Int. J. Hydrogen Energy 2019, 44, 23880–23898. [Google Scholar] [CrossRef]
  11. Kazi, M.-K.; Eljack, F.; El-Halwagi, M.M.; Haouari, M. Green Hydrogen for Industrial Sector Decarbonization: Costs and Impacts on Hydrogen Economy in Qatar. Comput. Chem. Eng. 2021, 145, 107144. [Google Scholar] [CrossRef]
  12. Chi, J.; Yu, H. Water Electrolysis Based on Renewable Energy for Hydrogen Production. Chin. J. Catal. 2018, 39, 390–394. [Google Scholar] [CrossRef]
  13. Shiva Kumar, S.; Himabindu, V. Hydrogen Production by PEM Water Electrolysis—A Review. Mater. Sci. Energy Technol. 2019, 2, 442–454. [Google Scholar] [CrossRef]
  14. Idriss, H. Hydrogen Production from Water: Past and Present. Curr. Opin. Chem. Eng. 2020, 29, 74–82. [Google Scholar] [CrossRef]
  15. Bhandari, R.; Trudewind, C.A.; Zapp, P. Life Cycle Assessment of Hydrogen Production via Electrolysis—A Review. J. Clean. Prod. 2014, 85, 151–163. [Google Scholar] [CrossRef]
  16. Ursua, A.; Gandia, L.M.; Sanchis, P. Hydrogen Production from Water Electrolysis: Current Status and Future Trends. Proc. IEEE 2012, 100, 410–426. [Google Scholar] [CrossRef]
  17. Hino, R.; Haga, K.; Aita, H.; Sekita, K. R&D on Hydrogen Production by High-Temperature Electrolysis of Steam. Nucl. Eng. Des. 2004, 233, 363–375. [Google Scholar] [CrossRef]
  18. Deng, Y.; Dewil, R.; Appels, L.; Li, S.; Baeyens, J.; Degrève, J.; Wang, G. Thermo-Chemical Water Splitting: Selection of Priority Reversible Redox Reactions by Multi-Attribute Decision Making. Renew. Energy 2021, 170, 800–810. [Google Scholar] [CrossRef]
  19. Nemitallah, M.A.; Alnazha, A.A.; Ahmed, U.; El-Adawy, M.; Habib, M.A. Review on Techno-Economics of Hydrogen Production Using Current and Emerging Processes: Status and Perspectives. Results Eng. 2024, 21, 101890. [Google Scholar] [CrossRef]
  20. Chao, C.; Deng, Y.; Dewil, R.; Baeyens, J.; Fan, X. Post-Combustion Carbon Capture. Renew. Sustain. Energy Rev. 2021, 138, 110490. [Google Scholar] [CrossRef]
  21. Licht, S. Over 18% Solar Energy Conversion to Generation of Hydrogen Fuel; Theory and Experiment for Efficient Solar Water Splitting. Int. J. Hydrogen Energy 2001, 26, 653–659. [Google Scholar] [CrossRef]
  22. Hara, D. Toward a Hydrogen Society—Introduction of Representative Projects in Japan. ECS Trans. 2019, 91, 3–7. [Google Scholar] [CrossRef]
  23. Arribas, L.; González-Aguilar, J.; Romero, M. Solar-Driven Thermochemical Water-Splitting by Cerium Oxide: Determination of Operational Conditions in a Directly Irradiated Fixed Bed Reactor. Energies 2018, 11, 2451. [Google Scholar] [CrossRef]
  24. Klahr, B.M.; Peterson, D.; Randolph, K.; Miller, E.L. Innovative Approaches to Addressing the Fundamental Materials Challenges in Advanced Water Splitting Technologies for Renewable Hydrogen Production. ECS Trans. 2017, 75, 3–11. [Google Scholar] [CrossRef]
  25. Herdem, M.S.; Sinaki, M.Y.; Farhad, S.; Hamdullahpur, F. An Overview of the Methanol Reforming Process: Comparison of Fuels, Catalysts, Reformers, and Systems. Int. J. Energy Res. 2019, 43, 5076–5105. [Google Scholar] [CrossRef]
  26. Moraes, T.S.; Cozendey da Silva, H.N.; Zotes, L.P.; Mattos, L.V.; Pizarro Borges, L.E.; Farrauto, R.; Noronha, F.B. A Techno-Economic Evaluation of the Hydrogen Production for Energy Generation Using an Ethanol Fuel Processor. Int. J. Hydrogen Energy 2019, 44, 21205–21219. [Google Scholar] [CrossRef]
  27. Liu, J.; Li, S.; Dewil, R.; Vanierschot, M.; Baeyens, J.; Deng, Y. Water Splitting by MnOx/Na2CO3 Reversible Redox Reactions. Sustainability 2022, 14, 7597. [Google Scholar] [CrossRef]
  28. Shaner, M.R.; Atwater, H.A.; Lewis, N.S.; McFarland, E.W. A Comparative Technoeconomic Analysis of Renewable Hydrogen Production Using Solar Energy. Energy Environ. Sci. 2016, 9, 2354–2371. [Google Scholar] [CrossRef]
  29. Kolb, G.J.; Diver, R.B.; Siegel, N. Central-Station Solar Hydrogen Power Plant. J. Sol. Energy Eng. 2007, 129, 179–183. [Google Scholar] [CrossRef]
  30. Zhang, Y.; Wang, Z.; Du, Z.; Li, Y.; Qian, M.; Van Herle, J.; Wang, L. Techno-Economic Analysis of Solar Hydrogen Production via PV Power/Concentrated Solar Heat Driven Solid Oxide Electrolysis with Electrical/Thermal Energy Storage. J. Energy Storage 2023, 72, 107986. [Google Scholar] [CrossRef]
  31. Xue, X.; Han, W.; Xin, Y.; Liu, C.; Jin, H.; Wang, X. Proposal and Energetic and Exergetic Evaluation of a Hydrogen Production System with Synergistic Conversion of Coal and Solar Energy. Energy 2023, 283, 128489. [Google Scholar] [CrossRef]
  32. Hosseini, S.E.; Wahid, M.A. Hydrogen from Solar Energy, a Clean Energy Carrier from a Sustainable Source of Energy. Int. J. Energy Res. 2020, 44, 4110–4131. [Google Scholar] [CrossRef]
  33. Charvin, P.; Stéphane, A.; Florent, L.; Gilles, F. Analysis of Solar Chemical Processes for Hydrogen Production from Water Splitting Thermochemical Cycles. Energy Convers. Manag. 2008, 49, 1547–1556. [Google Scholar] [CrossRef]
  34. Zoller, S.; Koepf, E.; Nizamian, D.; Stephan, M.; Patané, A.; Haueter, P.; Romero, M.; González-Aguilar, J.; Lieftink, D.; de Wit, E.; et al. A Solar Tower Fuel Plant for the Thermochemical Production of Kerosene from H2O and CO2. Joule 2022, 6, 1606–1616. [Google Scholar] [CrossRef]
  35. Khormi, N.A.; Fronk, B.M. Feasibility of High Temperature Concentrated Solar Power for Cogeneration of Electricity and Hydrogen Using Supercritical Carbon Dioxide Receiver Technology. In Proceedings of the ASME 2023 17th International Conference on Energy Sustainability, American Society of Mechanical Engineers, Washington, DC, USA, 10–12 July 2023. [Google Scholar]
  36. Lampe, J.; Menz, S. Optimized Operational Strategy of a Solar Reactor for Thermochemical Hydrogen Generation. Optim. Eng. 2024, 25, 29–61. [Google Scholar] [CrossRef]
  37. Dong, W.J.; Ye, Z.; Tang, S.; Navid, I.A.; Xiao, Y.; Zhang, B.; Pan, Y.; Mi, Z. Concentrated Solar Light Photoelectrochemical Water Splitting for Stable and High-Yield Hydrogen Production. Adv. Sci. 2024. early view. [Google Scholar] [CrossRef] [PubMed]
  38. Bentoumi, L.; Miles, A.; Korei, Z. Green Hydrogen Production by Integrating a Solar Power Plant with a Combined Cycle in the Desert Climate of Algeria. Sol. Energy 2024, 268, 112311. [Google Scholar] [CrossRef]
  39. Herrando, M.; Wang, K.; Huang, G.; Otanicar, T.; Mousa, O.B.; Agathokleous, R.A.; Ding, Y.; Kalogirou, S.; Ekins-Daukes, N.; Taylor, R.A.; et al. A Review of Solar Hybrid Photovoltaic-Thermal (PV-T) Collectors and Systems. Prog. Energy Combust. Sci. 2023, 97, 101072. [Google Scholar] [CrossRef]
  40. Benghanem, M.; Mellit, A.; Almohamadi, H.; Haddad, S.; Chettibi, N.; Alanazi, A.M.; Dasalla, D.; Alzahrani, A. Hydrogen Production Methods Based on Solar and Wind Energy: A Review. Energies 2023, 16, 757. [Google Scholar] [CrossRef]
  41. Liu, J.; Wang, J.; Tang, Y.; Jin, J.; Li, W. Solar Photovoltaic–Thermal Hydrogen Production System Based on Full-Spectrum Utilization. J. Clean. Prod. 2023, 430, 139340. [Google Scholar] [CrossRef]
  42. Boudries, R. Techno-Economic Study of Hydrogen Production Using CSP Technology. Int. J. Hydrogen Energy 2018, 43, 3406–3417. [Google Scholar] [CrossRef]
  43. IEA. The Future of Hydrogen. 2019. Available online: https://www.iea.org/reports/the-future-of-hydrogen (accessed on 1 February 2024).
  44. Deng, Y.; Li, S.; Appels, L.; Zhang, H.; Sweygers, N.; Baeyens, J.; Dewil, R. Steam Reforming of Ethanol by Non-Noble Metal Catalysts. Renew. Sustain. Energy Rev. 2023, 175, 113184. [Google Scholar] [CrossRef]
  45. Deng, Y.; Li, S.; Dewil, R.; Appels, L.; Yang, M.; Zhang, H.; Baeyens, J. Water Splitting by MnFe2O4/Na2CO3 Reversible Redox Reactions. RSC Adv. 2022, 12, 31392–31401. [Google Scholar] [CrossRef]
Figure 1. Current thermo-chemical H2 production cycles according to the types of Table 1. (a) metal-metal oxide; (b) metal oxide-metal oxide cycles; (c) metal oxide-metal hydroxide cycles; (d) metal oxide-metal sulfate cycles; (e) metal-metal halide cycles; (f) metal oxide-metal halide cycles.
Figure 1. Current thermo-chemical H2 production cycles according to the types of Table 1. (a) metal-metal oxide; (b) metal oxide-metal oxide cycles; (c) metal oxide-metal hydroxide cycles; (d) metal oxide-metal sulfate cycles; (e) metal-metal halide cycles; (f) metal oxide-metal halide cycles.
Sustainability 16 02883 g001
Figure 2. Solar-aided routes to produce H2, with (a) from water and fossil fuels; and (b) carbonaceous gases or solid feedstock.
Figure 2. Solar-aided routes to produce H2, with (a) from water and fossil fuels; and (b) carbonaceous gases or solid feedstock.
Sustainability 16 02883 g002
Figure 3. Expected H2 costs from hybrid solar PV and onshore wind systems. Reproduced by IEA; 2019; The Future of Hydrogen—Seizing today’s opportunities; https://www.iea.org/reports/the-future-of-hydrogen, accessed on 4 January 2024, License: [CC BY 4.0] [43].
Figure 3. Expected H2 costs from hybrid solar PV and onshore wind systems. Reproduced by IEA; 2019; The Future of Hydrogen—Seizing today’s opportunities; https://www.iea.org/reports/the-future-of-hydrogen, accessed on 4 January 2024, License: [CC BY 4.0] [43].
Sustainability 16 02883 g003
Figure 4. Temperature ranges and application areas of various CSP technologies.
Figure 4. Temperature ranges and application areas of various CSP technologies.
Sustainability 16 02883 g004
Figure 5. Experiments: (a) overall layout, (b) 26 mm I.D. vertical electrically heated reactor, (c) 9.6 mm I.D. horizontal electrically heated reactor, (d) fluidized bed reactor (50 mm I.D.) in the solar cavity, and (e) illustration of the reactor depicted in (d).
Figure 5. Experiments: (a) overall layout, (b) 26 mm I.D. vertical electrically heated reactor, (c) 9.6 mm I.D. horizontal electrically heated reactor, (d) fluidized bed reactor (50 mm I.D.) in the solar cavity, and (e) illustration of the reactor depicted in (d).
Sustainability 16 02883 g005
Figure 6. Methane conversion for different residence times at 650 and 700 °C.
Figure 6. Methane conversion for different residence times at 650 and 700 °C.
Sustainability 16 02883 g006
Figure 7. Conversion of methanol with Co/α-Al2O3 (a) and MnFe2O4 (b) catalysts.
Figure 7. Conversion of methanol with Co/α-Al2O3 (a) and MnFe2O4 (b) catalysts.
Sustainability 16 02883 g007
Figure 8. Effect of residence time and temperature on the conversion of NH3.
Figure 8. Effect of residence time and temperature on the conversion of NH3.
Sustainability 16 02883 g008
Figure 9. Results for the 9% Co/α-Al2O3 and 4.5% Co-4.5%Ni/α-Al2O3 catalysts (a,b).
Figure 9. Results for the 9% Co/α-Al2O3 and 4.5% Co-4.5%Ni/α-Al2O3 catalysts (a,b).
Sustainability 16 02883 g009
Figure 10. H2 yields using the MnOx/Na2CO3 cycle in the solar furnace.
Figure 10. H2 yields using the MnOx/Na2CO3 cycle in the solar furnace.
Sustainability 16 02883 g010
Figure 11. Redox cycle experiments (reproduced from [45], courtesy of RSC Advances). (a) Weight changes in the oxidation (3 h) and reduction cycles (4.5 h), respectively. (b) Weight changes in the oxidation (3 h) and reduction cycles (6 h), respectively.
Figure 11. Redox cycle experiments (reproduced from [45], courtesy of RSC Advances). (a) Weight changes in the oxidation (3 h) and reduction cycles (4.5 h), respectively. (b) Weight changes in the oxidation (3 h) and reduction cycles (6 h), respectively.
Sustainability 16 02883 g011
Figure 12. Layout of the experimental set-up. 1. Solar Cavity Receiver 1-a (with associated heliostats 1-b) and either vertical or horizontal fluidized bed particle carrier heating. 2. Hot particle carrier storage with possibly supplementary heating from PV or wind turbine (or waste heat). 3. Free-bearing screw conveyors. 4. Moving-bed indirect conversion reactor (multi-tubular catalyst bed). 5. Recycling of carrier particles to 1.
Figure 12. Layout of the experimental set-up. 1. Solar Cavity Receiver 1-a (with associated heliostats 1-b) and either vertical or horizontal fluidized bed particle carrier heating. 2. Hot particle carrier storage with possibly supplementary heating from PV or wind turbine (or waste heat). 3. Free-bearing screw conveyors. 4. Moving-bed indirect conversion reactor (multi-tubular catalyst bed). 5. Recycling of carrier particles to 1.
Sustainability 16 02883 g012
Figure 13. Purification cycles of H2.
Figure 13. Purification cycles of H2.
Sustainability 16 02883 g013
Figure 14. Solar H2 production using the MnFe2O4/Na2CO3 water splitting cycle at 735 °C (■) and 710 °C (●).
Figure 14. Solar H2 production using the MnFe2O4/Na2CO3 water splitting cycle at 735 °C (■) and 710 °C (●).
Sustainability 16 02883 g014
Figure 15. Novel concentrated solar dish technology developed by the Chinese Academy of Sciences.
Figure 15. Novel concentrated solar dish technology developed by the Chinese Academy of Sciences.
Sustainability 16 02883 g015
Table 1. Characteristics of H2 generation methods.
Table 1. Characteristics of H2 generation methods.
H2 Production MethodsH2 Cost
USD/kg H2
AdvantagesDrawbacks
Auto-thermal reforming1.5
Low operating temperature.
Limited methane slip.
Air is required.
O2/CO2—byproducts.
Partial oxidation1.5
No catalyst needed.
Requires a lower degree of feedstock desulfurization.
Limited methane slip.
Low H2/CO ratio.
Soot formation.
High operating temperature.
Steam reforming3–5
Mature technology.
No O2 needed.
Lower operating temperature.
Highest H2/CO ratio.
High production of CO.
High production of CO2.
Gasification/pyrolysis1.6–2.1
Nearly CO2-neutral.
Abundant.
Cheap biomass feedstock.
Variable H2 yield.
Formation volumes.
Electrolysis5–7
Zero emission.
O2—byproduct.
Existing infrastructure.
Currently uses fossil fuel-based electricity.
Water thermolysis8
Clean and sustainable.
O2—byproduct.
Abundant feedstock.
Elements toxicity.
Corrosion problems.
High capital costs.
Bio photolysis1
CO2 consumed.
O2—byproduct.
Mild operation conditions.
Lower H2 yields.
O2 sensitivity.
Sunlight needed.
Requires large reactor volumes.
Photolysis8–10
Abundant feedstock.
O2—byproduct.
No emission.
Low efficiency.
Requires sunlight.
Table 3. Thermo-chemical redox H2 cycles.
Table 3. Thermo-chemical redox H2 cycles.
TypesNumber of Steps in the Complete Cycle
Type 1: Metal–metal oxide T w o
Type 2: Metal oxide–metal oxide F o u r
Type 3: Metal oxide–metal hydroxide T h r e e
Type 4: Metal oxide–metal sulfate T h r e e
Type 5: Metal–metal halide T h r e e
Type 6: Metal oxide–metal halides F o u r
Table 5. Carbon selectivity of steam methanol reforming.
Table 5. Carbon selectivity of steam methanol reforming.
MnFe2O4Co-Al2O3
Selectivity (%)573 K673 K573 K
CO38.1227.8529.24
CH45.7620.5425.88
CO256.1151.6144.88
Table 6. Gas hourly space velocities (GHSVs) of different hydrogen production systems.
Table 6. Gas hourly space velocities (GHSVs) of different hydrogen production systems.
SystemGHSV
NH310 L H2 h−1 kgcat−1
CH3OH900 L H2 h−1 kgcat−1
C2H5OH13,000 L H2 h−1 kgcat−1
Table 7. Comparing H2 production methods (from 2017 to 2023).
Table 7. Comparing H2 production methods (from 2017 to 2023).
LCOH (USD/kg H2)GHG Emission
(kg CO2-eq/kg H2)
Methane steam reforming1.258.5–11
Coal gasification1.512.7–16.8
PEM electrolysis (renewable energy)7.70
Biomass gasification3 (for large scale)
>5 (otherwise)
0 (if biomass is carbon neutral)
Pyrolysis1.8–2.1-
Methanol steam reforming2.2>3
Ethanol steam reforming1–2.5>3
Methane decomposition0.4–1.5 *0.85–1.05
Ammonia decomposition1.5–2.50
Redox water splitting~1.00
* Carbon black is sold at EUR 470 and EUR 1870 per ton.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Deng, Y.; Li, S.; Liu, H.; Zhang, H.; Baeyens, J. Recent Research in Solar-Driven Hydrogen Production. Sustainability 2024, 16, 2883. https://doi.org/10.3390/su16072883

AMA Style

Deng Y, Li S, Liu H, Zhang H, Baeyens J. Recent Research in Solar-Driven Hydrogen Production. Sustainability. 2024; 16(7):2883. https://doi.org/10.3390/su16072883

Chicago/Turabian Style

Deng, Yimin, Shuo Li, Helei Liu, Huili Zhang, and Jan Baeyens. 2024. "Recent Research in Solar-Driven Hydrogen Production" Sustainability 16, no. 7: 2883. https://doi.org/10.3390/su16072883

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