Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 3160 2023-12-18 10:16:35 |
2 references update Meta information modification 3160 2023-12-19 06:26:38 | |
3 format layout + 2 word(s) 3162 2023-12-22 06:58:19 | |
4 layout Meta information modification 3162 2023-12-22 06:59:27 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Di Nardo, A.; Calabrese, M.; Venezia, V.; Portarapillo, M.; Turco, M.; Di Benedetto, A.; Luciani, G. Hydrogen Production from Bio-Alcohols. Encyclopedia. Available online: https://encyclopedia.pub/entry/52860 (accessed on 19 May 2024).
Di Nardo A, Calabrese M, Venezia V, Portarapillo M, Turco M, Di Benedetto A, et al. Hydrogen Production from Bio-Alcohols. Encyclopedia. Available at: https://encyclopedia.pub/entry/52860. Accessed May 19, 2024.
Di Nardo, Alessandra, Marcella Calabrese, Virginia Venezia, Maria Portarapillo, Maria Turco, Almerinda Di Benedetto, Giuseppina Luciani. "Hydrogen Production from Bio-Alcohols" Encyclopedia, https://encyclopedia.pub/entry/52860 (accessed May 19, 2024).
Di Nardo, A., Calabrese, M., Venezia, V., Portarapillo, M., Turco, M., Di Benedetto, A., & Luciani, G. (2023, December 18). Hydrogen Production from Bio-Alcohols. In Encyclopedia. https://encyclopedia.pub/entry/52860
Di Nardo, Alessandra, et al. "Hydrogen Production from Bio-Alcohols." Encyclopedia. Web. 18 December, 2023.
Hydrogen Production from Bio-Alcohols
Edit

Energy and environmental issues are of great importance in the present era. The transition to renewable energy sources necessitates technological, political, and behavioral transformations. Hydrogen is a promising solution, and many countries are investing in the hydrogen economy. The incorporation of hydrogen for efficient energy transport and storage and its integration into the transport sector are crucial measures. However, to fully develop a hydrogen-based economy, the sustainability and safety of hydrogen in all its applications must be ensured. 

hydrogen production bio-alcohol

1. Introduction

Population growth as well as industrialization have been accompanied by a huge increase in energy consumption, causing severe environmental impact in terms of fossil fuel depletion, pollution, and increased content of greenhouse gases [1]. The grey range with its central black line draws a corresponding estimate of the human-caused share of historical warming. Coloured areas show the global surface temperature projections, and thick coloured central lines show the median estimate as a function of cumulative CO2 emissions for different scenarios, ranging from very low (SSP1-1.9), low (SSP1-2.6), intermediate (SSP2-4.5), high (SSP3-7.0), to very high (SSP5-8.5) [2].
To limit human-caused global warming to a specific level, we need to restrain the cumulative CO2 emissions, reaching net zero or net negative CO2 emissions, along with reductions of other GHG emissions. To this end in the Paris Agreement countries agreed to reach the following targets: a reduction of 50–52% of CO2 emissions by 2030; carbon pollution-free electricity by 2035; net ZERO GHG emissions by 2050 [3]
To reach these goals, a complete energetic transition towards renewable, cleaner, and more sustainable sources is required, with the most important strategy being the development of H2 technologies and their widespread use in human activities. H2 is intrinsically clean since it only produces water, thus contributing to decarbonization. H2 can be produced by different means, including renewable sources. It releases a significant amount of heat when burned without pollution. Furthermore, it can be exploited in a huge range of applications, replacing traditional fossil fuels and including the transport, heat, industry, and electricity sectors [4]. Finally, because of the plethora of production technologies, its availability is not affected by outside influences [5]. Despite the significant technological potential and the possibility of transitioning to a low-carbon economy, hydrogen as an energy carrier remains limited, with only 6% of H2 production being used for this purpose. Most of the hydrogen is instead used as an intermediate in industrial chemical production, emphasizing its constrained role in meeting global energy demands [6].
To reach the goal of widespread use of H2, all the related technologies must be developed and optimized to meet sustainable targets. Figure 1 shows the H2 cycle, evidencing that it is based on four cornerstones: production, storage, transportation, and use. Each step must be managed to eventually produce clean energy; H2 must be produced, and then the first action is the optimization of H2 production unit. Produced H2 must be stored and transported in a green, sustainable, and safe way. Finally, H2 should be used for clean and efficient energy production through high-performance devices that ensure significant energy conversion rates by using fuel cells.
Figure 1. The H2 cycle.

2. Hydrogen Production from Bio-Alcohols: Focus on Bioethanol

Biowaste-derived alcohols represent an important alternative feedstock for carbon-neutral hydrogen production. Since they can be produced directly from renewable biomass, they have a sustainable feature [7]. CO2 emissions from producing hydrogen using bioderived sources are offset by the CO2 previously absorbed by these sources, resulting in net zero CO2 emissions [8].
Bioethanol is traditionally made from plant by-products such as sugarcane, corn, and wheat (first-generation). However, lignocellulosic (second-generation) and algal biomass (third-generation) are now explored due to food sustainability concerns [9]. However, second-generation raw materials necessitate further procedures, including pre-treating lignocellulose, to generate fermentable substances. The challenge remains in maintaining the cost-effectiveness of these processes. Additionally, research in the third-generation sector is still in its early stages due to limited investment and design hurdles [10]. Annual production of bioethanol is steadily increasing, with a projected global output and consumption of nearly 134.5 billion liters by 2024. The cost of raw materials has a significant influence on bioethanol production expenses, accounting for between 40 and 75% of the total cost [11]. At current pricing, the anticipated cost of advanced bioethanol is between USD 1.04 and USD 1.45 per liter of gasoline equivalent (LGE) [12]. Minimising greenhouse gas emissions, ensuring energy security, and promoting agricultural practices in rural areas are all measures that can assist in offsetting the production of bioethanol.
Researchers have investigated a variety of methods to manufacture hydrogen from light alcohols, such as steam reforming (SR), partial oxidation (POx), autothermal reforming (ATR), photocatalysis (PC), and electrocatalysis (EC). 

2.1. Reforming

Steam reforming (SR) is a highly advanced technique used in the petrochemical industry to convert natural gas, primarily methane, into hydrogen. However, it results in 8–13 Kg CO2eq/KgH2 emissions, both as a product reaction and for heating the reactor due to the reaction endothermicity [13]. Today, the research aims to expand the capabilities and adaptability of SR technology beyond its dependence on natural gas. One approach involves investigating alternative reactants, such as oxygenated hydrocarbons such as methanol [14][15][16], ethanol [14][17][18][19], glycerol [20][21], dimethyl ether, acetone, and acetic acid, as well as heavier hydrocarbons ranging from C3 to C10 [22].
For oxygenated hydrocarbons, the stoichiometric SR reaction (1) is as follows: The CO can be further transformed to produce additional hydrogen and carbon dioxide through the exothermic water-gas shift (WGS) reaction (2).
C x H y O z + ( 2 x z )   H 2 O     x   C O + 2 ( 2 x     z ) +   y 2   H 2
C O + H 2 O     C 2 O + H 2
Another reforming technology is partial oxidation (POx) (3), in which partial combustion of the fuel produces hydrogen.
C x H y O z + ( 2 x     z ) 2   O 2     x   C O 2 + y 2   H 2

During POx, the fuel reacts with a limited supply of air or pure oxygen. The exothermic nature of the reaction means that it releases heat, avoiding the need for external heating. It is important to control the amount of air or oxygen supplied during partial oxidation to achieve the desired products and avoid complete oxidation, which would result in a reduced amount of hydrogen produced [23].

Autothermal reforming (ATR) (4) is a combination of both partial oxidation (exothermic) and steam reforming processes (endothermic) to supply the required heat and enable self-sustained reforming.
C x H y O z + [ 2 x     ( z + 1 ) ]   H 2 O + 1 2   O 2     x   C O 2 + 2   [   2 x     ( z   + 1 ) ] +   y 2   H 2
Fuel is first oxidized to heat up the reactor and then reacts with steam to produce hydrogen-rich syngas [24]. ATR has the potential to achieve high hydrogen yield and hydrogen selectivity because syngas can be produced from air or H2O. By using the POx to produce heat and the SR to increase hydrogen production, the ATR technique produces a thermally neutral process [25].
In the case of ethanol, ethanol SR (ESR), POx, and ATR reactions are reported in the following (5)–(7) and are accompanied by the water-gas shift reaction (WGS) (2) to produce hydrogen:
C2H5OH + H2O → 2 CO + 4 H2                   
C2H5OH + ½ O2 → CO + 3 H2                     
C2H5OH + 2 H2O + ½ O2 → 2 CO2 + 5 H2           
In Table 1 are reported the operating conditions (temperature, pressure, feed ratio) of the ethanol reforming technologies.
The reaction pathway of ESR has been extensively studied, and the mechanism complexity is widely reported [26][27][28] since different parallel reactions occur simultaneously with hydrogen production. The product distribution depends on both the nature of the catalyst (i.e., active sites, support, precursor, preparation methods) and the reaction conditions, e.g., temperature, water/ethanol ratio, and feed flow rate. The complex set of by-products includes carbon monoxide (CO), methane (CH4), ethane (C2H6), ethylene (C2H4), and acetaldehyde (CH3CHO) [29].
Supported metal catalysts are commonly employed for ethanol steam reforming (ESR), in which the support plays a crucial role in activating steam to produce oxygen. Oxygen stabilizes the metal particles present in the catalyst and promotes the gasification of coke, while the metal phases break the hydrocarbon bonds [7]. Non-noble metal-based catalysts are widely employed for ethanol reforming due to their relatively low cost and good activity for C-C and C-H bond cleavage [30]. Among Ni, Co, Fe, and Cu species, the highest activity and selectivity for hydrogen and the lowest formation of CH4 are achieved by Co [31][32]. Noble metal-based catalysts (Pt, Pd, Rh, and Ru) have high activity and stability in ESR, but the resource scarcity and relatively high costs represent an obstacle for their large-scale application. In general, Rh-based catalysts emerge as the most active and selective for H2 production, even at low temperatures [33][34][35]. The reaction is highly influenced by the kind of support: basic solids encourage the dehydrogenation to acetaldehyde, while acidic ones cause the dehydration to ethylene, a key precursor to coke in ESR [36]. The most investigated and widely adopted support is Al2O3, but due to its acidic nature, it induces coke formation. The selectivity of H2 for Co-based catalysts decreased in the following order: Co/Al2O3 > Co/ZrO2 > Co/MgO > Co/SiO2 > Co/C [37], while the higher activity and hydrogen yield are achieved by Co/ZrO2 [38]. The Rh/Al2O3 catalyst showed the highest selectivity for H2, while Rh/CeO2–ZrO2 exhibited the highest yield in H2 [39].
For the ethanol POx, Salge et al. [40] reported that the H2 yield increases in the order Rh–Ru > Rh > Pd > Pt for an alumina-supported catalyst and that Rh/CeO2 appears as the most stable catalyst with a H2 selectivity, thanks to the promotion of WGS and the limitation of ethanol dehydration by ceria [41].
Also, in the case of ethanol ATR, Youn et al. [42] showed that nickel catalysts supported by the basic supports, ZrO2 and TiO2, had improved catalytic performance (hydrogen yield) and resistance to coke formation. Moreover, Gutierrez et al. [43] demonstrated that the Rh-containing catalysts are more suitable than the commercial Ni/Al2O3 catalysts for the ATR of ethanol.
Table 1. Operating conditions (temperature, pressure, feed ratio) of the ethanol reforming technologies for hydrogen production.
Technology Temperature Pressure Feed Ratio Ref.
Steam Reforming 450–650 °C 2–25 bar Water/Ethanol = 3–6 [7][22]
Partial Oxidation 200–400 °C 1–10 bar Air/Ethanol = 4 [41][44]
Autothermal Reforming 600–900 °C 1–10 bar Ethanol/Air/Water = 1/0.3–0.5/2–3 [41][45]

2.2. Photocatalysis

Photo-assisted hydrogen production processes, a promising way to utilize solar energy and water or other renewable feedstocks, have received significant attention over the last few years [46]. Photocatalysis is a process in which light is absorbed by a semiconducting material to generate electron-hole pairs. When the energy of the irradiated light photon is greater than the band-gap energy of the semiconductor, the electrons migrate to the conduction band (CB) with positive holes left on the valence band (VB). Holes can oxidize either water or organic/inorganic compounds and split them into H+ protons on VB; electrons reduce the H+ species to produce H2 on CB (Figure 2). However, the charge carrier reaction with the target substance competes with the recombination of an electron and a hole [47][48]. If bio-available oxygenated compounds are employed as sacrificial agents, this approach can be seen as nearly carbon-neutral since the CO2 generated can be transformed back into biomass via plant photosynthesis [48][49]. A variety of organic compounds have been investigated, such as methanol [50][51], ethanol [52], glycerol [53][54], and raw biomass [55]. Compared to pure water splitting, the lower Gibbs free energy change leads to more efficient hydrogen production [56], and the formation of CO2 as the product of the oxidation half reaction instead of O2 decreases the recombination reaction [57]. The fundamental reactions are reported in (8)–(9) and in Figure 2.
CxHyOz + (2x − z) H2O → (4x − 2z + y) H+ + x CO2    
2 H+ + 2 e → H2     
Figure 2. Schematic representation of the photocatalytic process of bio-alcohols and oxygenated hydrocarbons.
For ethanol photoreforming, several mechanisms [56] were proposed. Among those, the most reliable considers the direct oxidation by photoexcited holes in acetaldehyde (10) and the simultaneous reduction of the H+ to H2 (9). The acetaldehyde is then converted to CH4 and CO. Through the WSR reaction (7), CO and CH4 produce CO2 and other H2 [58][59][60].
C2H5OH + 2 H+ → CH3CHO + 2 H+     
TiO2-based photocatalysts have received the most attention among all photocatalysts, largely because of their favored characteristics, including an energy band location that can drive a variety of redox processes and, more importantly, H+ reduction, good chemical stability, and low toxicity [61]. However, the low efficiency for hydrogen production is due to fast electron-hole recombination in the bulk or on the surface of semiconductor particles. A promising approach to avoiding that is the deposition of metal nanoparticles on the surface of a photocatalyst that tend to take the electron and avoids recombination [62]. Noble metals including Au, Pd, and Pt are preferred for this purpose, even though Cu species have been proposed as a more economic choice [63][64][65][66]. Other key factors that influence hydrogen production are solution pH, alcohol concentration, temperature, and catalyst load [67]. Increasing the alcohol concentration initially increases the rate, but excessive amounts hinder water adsorption and decrease photocatalytic activity. The same rate trend occurs in relation to the catalyst load: at low catalyst loads, photon adsorption is not complete, while high catalyst loads lead to non-uniform adsorption of light photons. The solution pH affects the reaction rate due to the agglomeration of photocatalyst particles and to variation of the surface charge, which influence the absorption of H+ and alcohols [68]. Estahbanati et al. [58] reported that the maximum amount of hydrogen produced is around a pH of 4 in the presence of Pt/TiO2 for the use of ethanol, methanol, and glycerol as sacrificial agents. The thermal activation energies in these systems are typically less than 30 kJ mol−1, indicating milder activation steps compared to conventional catalytic reactions, such as product desorption and photoexcited carrier migration.

2.3. Electrolysis

Nowadays, water electrolysis represents the most developed sustainable and clean hydrogen production process, but it covers only 4% of industrial hydrogen production processes globally due to economic issues [69]. In fact, even when utilizing noble metal-based catalysts (Pt, Ru, Ir), the low kinetics of water oxidation, which implies high cell voltage (i.e., high energy consumption), makes this technique expensive and not competitive with the main hydrogen production method from natural gas [70]. Therefore, using bio-alcohols as hydrogen carriers in an electrolysis cell can be an interesting alternative, since their reversible oxidation potentials are much lower than that of water (ca. 0.1 V against 1.23 V under standard conditions), so the cell voltages or hydrogen production are lower (and so is the energy consumption) [71]. Methanol [72], ethanol [73], glycerol [74], and ethylene glycol [75] have been considered for hydrogen production through their electrolysis in proton exchange membrane electrolysis cells (PEMECs).
Figure 3 shows the principal of electrochemical decomposition of a bio-alcohol on a PEMEC. Taking ethanol as an example, it is fed to the anode and completely oxidized in the presence of water, releasing CO2 and protons (H+) (11). The protons that reach the cathode compartment, after crossing over the membrane, are reduced to H2 (12).
C2H5OH + 3 H2O → 2 CO2 + 12 H+ + 12 e−           
12 H+ + 12 e → 6 H2          
Figure 3. Schematic representation of the electrolysis process of bio-alcohols and oxygenated hydrocarbons.
This corresponds to the electrochemical reforming of ethanol into hydrogen and carbon dioxide, according to the overall reaction:
C2H5OH + 3 H2O → 6 H2 + 2 CO2        
Both water and ethanol electrolysis require external energy; however, the energy needed for ethanol is less than that for water (ΔH = +58 kJ/mole H2 for ethanol vs. ΔH = +286 kJ/mole H2 for water) [71]. The lowering of cell voltage is required to boost the energy efficiency of an alcohol PEMEC. The decrease in cell voltage corresponds directly to the increase in the rate of electrochemical reactions happening at both electrodes, with a particular emphasis on the anode, where alcohol oxidation takes place [70]. Unfortunately, the electrooxidation of ethanol is a slow process with overpotentials of 0.3–0.4 V. Furthermore, platinum is known to be rapidly poisoned on its surface by adsorbed species (COads) formed by alcohol dissociative adsorption [76]. To prevent this phenomenon, one approach is to modify the electrode by introducing oxygenated species (e.g., OHads) that are generated through the dissociation of water at lower potentials. This modification facilitates the complete oxidation of COads species to carbon dioxide [77]. It has generally been found that combining Pt with other metals, such as Ni, Ru [73][78], Rh [70][71], Co [79], and Sn [80][81][82][83], in bi- and tri-metallic [84][85] catalysts increase the electrochemical performance of the anode, lowering the rate of catalyst deactivation. Nickel-containing Pt-based catalysts have demonstrated better electrochemical performance compared to other well-studied promoters such as Ru, especially at high potentials, while also being cost-effective [86]. A Pt-Ni-based anodic catalyst supported by graphene nanopatterns has been recently developed by Serrano-Jiménez et al. [87], achieving competitive current density and energy consumption.
As reported above, if ethanol is produced from renewable sources or organic waste, the process can contribute to a carbon-neutral balance. Moreover, bioethanol is relatively easy to obtain from renewable sources and thus has abundant availability. Table 2 shows the advantages and disadvantages, costs, and efficiency of the different investigated methods for hydrogen production from bio-ethanol.
Table 2. Pros and cons, costs, and yield for hydrogen production by ethanol.
H2 Production
Process from Ethanol
Advantages Disadvantages Performances Ref.
Steam Reforming
Well-established and extensively used
Large-scale applicability
Higher selectivity compared to POx and ATR
Coke formation and deactivation of catalyst
High heat requirements
H2 Yield: 60–95%
EtOH Conversion:
90–100%
H2 Selectivity: 70–90%
@Ni/Al2O3,
250–500 °C,
H2O/EtOH = 3–6
H2 production cost:
1.58–2.6 USD/kgH2
[18][44][88][89]
Partial Oxidation
Temperature lower than SR and ATR
Quick response time and fast start-up
Lower ratio of H2/CO
Higher reaction temperature than ESR
H2 Selectivity: 60–97%
EtOH Conversion:
@Noble metal/Al2O3,
500–600 °C,
H2O/O2/EtOH = 3/0.3/1
[90][91]
Autothermal
Reforming
Thermal neutrality
High H2 selectivity
No external heat supply
Complexity process
High cost
Higher reaction temperature than ESR
H2 Yield: 47–94%
EtOH Conversion:
84–97%
@Rh/Al2O3
H2 Selectivity: 40–60%
@Pt/Al2O3
400–600 °C,
H2O/O2/EtOH = 3/0.3/1
[92][93]
Photocatalysis
Renewable primary energy source
High energy efficiency
Applicability at a small scale
Expensive catalysts
Sensitivity to environmental condition
Limited hydrogen production rate
H2 production rate:
30–34 mmol/gh
@Pt/TiO2
[94][95]
Electrocatalysis
Scale flexibility
Co-generation of electricity and hydrogen
Expensive catalysts
Energy losses
Need for continuous and constant electricity
H2 volume: 102 mL
@Pt/C,
20 °C
j = 50 mA/cm−2
[71]
The production of hydrogen from ethanol entails several distinct technologies, each with their own advantages and challenges. The decision on which method to utilize depends on factors such as scalability, economics, and efficiency. Therefore, it is essential to continue research and development in order to overcome limitations. Sustainable hydrogen production is of utmost importance. Progress in catalyst development, process optimization, and cost reduction may revolutionize ethanol-based hydrogen production, making it economically feasible and environmentally sustainable.
In this scenario, researchers propose a novel hydrogen production process, recently patented [96], so-called cyan hydrogen, from bio-alcohols and sodium metaborate, a by-product of little industrial interest, as a valid alternative. This name is inspired by a combination of the green and blue processes, due to the key role played by water and the low carbon content in the gas phase, respectively. The developed process is constituted by a sequential discontinuous step in which bio-alcohols and water are alternatively fed to the sodium metaborate. Recently, researchers demonstrated the feasibility of the process in the presence of ethanol [97]. The process results in the simultaneous production of a hydrogen-rich stream (98% v/v) and a polymeric compound with a repetitive carbon pattern, −CH2−CH2−, i.e., polyethylene structure.

References

  1. Fayyazbakhsh, A.; Bell, M.L.; Zhu, X.; Mei, X.; Koutný, M.; Hajinajaf, N.; Zhang, Y. Engine Emissions with Air Pollutants and Greenhouse Gases and Their Control Technologies. J. Clean. Prod. 2022, 376, 134260.
  2. Calvin, K.; Dasgupta, D.; Krinner, G.; Mukherji, A.; Thorne, P.W.; Trisos, C.; Romero, J.; Aldunce, P.; Barrett, K.; Blanco, G.; et al. IPCC, 2023: Climate Change 2023: Synthesis Report. Contribution of Working Groups I, II and III to the Sixth Assessment Report. of the Intergovernmental Panel on Climate Change; Arias, P., Bustamante, M., Elgizouli, I., Flato, G., Howden, M., Méndez-Vallejo, C., Pereira, J.J., Pichs-Madruga, R., Rose, S.K., Saheb, Y., et al., Eds.; IPCC: Geneva, Switzerland, 2023.
  3. Bouckaert, S.; Pales, A.F.; McGlade, C.; Remme, U.; Wanner, B.; Varro, L.; D’Ambrosio, D.; Spencer, T. Net Zero by 2050—A Roadmap for the Global Energy Sector; International Energy Agency: Paris, France, 2021; Available online: https://www.iea.org/reports/net-zero-by-2050 (accessed on 18 November 2023).
  4. Schmidt, J.; Gruber, K.; Klingler, M.; Klöckl, C.; Ramirez Camargo, L.; Regner, P.; Turkovska, O.; Wehrle, S.; Wetterlund, E. A New Perspective on Global Renewable Energy Systems: Why Trade in Energy Carriers Matters. Energy Environ. Sci. 2019, 12, 2022–2029.
  5. Ahmed, M.R.; Barua, T.; Das, B.K. A Comprehensive Review on Techno-Environmental Analysis of State-of-the-Art Production and Storage of Hydrogen Energy: Challenges and Way Forward. Energy Sources Part A Recovery Util. Environ. Eff. 2023, 45, 5905–5937.
  6. Eljack, F.; Kazi, M.K. Prospects and Challenges of Green Hydrogen Economy via Multi-Sector Global Symbiosis in Qatar. Front. Sustain. 2020, 1, 612762.
  7. Xiang, H.; Xin, R.; Prasongthum, N.; Natewong, P.; Sooknoi, T.; Wang, J.; Reubroycharoen, P.; Fan, X. Catalytic Conversion of Bioethanol to Value-Added Chemicals and Fuels: A Review. Resour. Chem. Mater. 2022, 1, 47–68.
  8. Culaba, A.B.; Mayol, A.P.; San Juan, J.L.G.; Ubando, A.T.; Bandala, A.A.; Concepcion, R.S.; Alipio, M.; Chen, W.H.; Show, P.L.; Chang, J.S. Design of Biorefineries towards Carbon Neutrality: A Critical Review. Bioresour. Technol. 2023, 369, 128256.
  9. Tse, T.J.; Wiens, D.J.; Reaney, M.J.T. Production of Bioethanol—A Review of Factors Affecting Ethanol Yield. Fermentation 2021, 7, 268.
  10. Melendez, J.R.; Mátyás, B.; Hena, S.; Lowy, D.A.; El Salous, A. Perspectives in the Production of Bioethanol: A Review of Sustainable Methods, Technologies, and Bioprocesses. Renew. Sustain. Energy Rev. 2022, 160, 112260.
  11. Bušić, A.; Mardetko, N.; Kundas, S.; Morzak, G.; Belskaya, H.; Šantek, M.I.; Komes, D.; Novak, S.; Šantek, B. Bioethanol Production from Renewable Raw Materials and Its Separation and Purification: A Review. Food Technol. Biotechnol. 2018, 56, 289–311.
  12. IRENA. Available online: https://www.irena.org/Energy-Transition/Technology/Transportation-costs/Bioethanol (accessed on 15 November 2023).
  13. Suer, J.; Traverso, M.; Jäger, N. Carbon Footprint Assessment of Hydrogen and Steel. Energies 2022, 15, 9468.
  14. Bepari, S.; Kuila, D. Steam Reforming of Methanol, Ethanol and Glycerol over Nickel-Based Catalysts—A Review. Int. J. Hydrogen Energy 2020, 45, 18090–18113.
  15. Kappis, K.; Papavasiliou, J.; Avgouropoulos, G. Methanol Reforming Processes for Fuel Cell Applications. Energies 2021, 14, 8442.
  16. Xu, X.; Shuai, K.; Xu, B. Review on Copper and Palladium Based Catalysts for Methanol Steam Reforming to Produce Hydrogen. Catalysts 2017, 7, 183.
  17. Ogo, S.; Sekine, Y. Recent Progress in Ethanol Steam Reforming Using Non-Noble Transition Metal Catalysts: A Review. Fuel Process. Technol. 2020, 199, 106238.
  18. Anil, S.; Indraja, S.; Singh, R.; Appari, S.; Roy, B. A Review on Ethanol Steam Reforming for Hydrogen Production over Ni/Al2O3 and Ni/CeO2 Based Catalyst Powders. Int. J. Hydrogen Energy 2022, 47, 8177–8213.
  19. Shtyka, O.; Dimitrova, Z.; Ciesielski, R.; Kedziora, A.; Mitukiewicz, G.; Leyko, J.; Maniukewicz, W.; Czylkowska, A.; Maniecki, T. Steam Reforming of Ethanol for Hydrogen Production: Influence of Catalyst Composition (Ni/Al2O3, Ni/Al2O3–CeO2, Ni/Al2O3–ZnO) and Process Conditions. React. Kinet. Mech. Catal. 2021, 132, 907–919.
  20. Adeniyi, A.G.; Ighalo, J.O. A Review of Steam Reforming of Glycerol. Chem. Pap. 2019, 73, 2619–2635.
  21. Fasolini, A.; Cespi, D.; Tabanelli, T.; Cucciniello, R.; Cavani, F. Hydrogen from Renewables: A Case Study of Glycerol Reforming. Catalysts 2019, 9, 722.
  22. Azizan, M.T.; Aqsha, A.; Ameen, M.; Syuhada, A.; Klaus, H.; Abidin, S.Z.; Sher, F. Catalytic Reforming of Oxygenated Hydrocarbons for the Hydrogen Production: An Outlook. Biomass Convers. Biorefinery 2020, 13, 8441–8464.
  23. Makaryan, I.A.; Salgansky, E.A.; Arutyunov, V.S.; Sedov, I.V. Non-Catalytic Partial Oxidation of Hydrocarbon Gases to Syngas and Hydrogen: A Systematic Review. Energies 2023, 16, 2916.
  24. Balopi, B.; Moyo, M.; Gorimbo, J. Autothermal Reforming of Bio-Ethanol: A Short Review of Strategies Used to Synthesize Coke-Resistant Nickel-Based Catalysts. Catal. Lett. 2022, 152, 3004–3016.
  25. Baruah, R.; Dixit, M.; Basarkar, P.; Parikh, D.; Bhargav, A. Advances in Ethanol Autothermal Reforming. Renew. Sustain. Energy Rev. 2015, 51, 1345–1353.
  26. Wang, W.; Wang, Y.Q. Thermodynamic Analysis of Steam Reforming of Ethanol for Hydrogen Generation. Int. J. Energy Res. 2008, 32, 1432–1443.
  27. Haryanto, A.; Fernando, S.; Murali, N.; Adhikari, S. Current Status of Hydrogen Production Techniques by Steam Reforming of Ethanol: A Review. Energy Fuels 2005, 19, 2098–2106.
  28. Vicente, J.; Ereña, J.; Montero, C.; Azkoiti, M.J.; Bilbao, J.; Gayubo, A.G. Reaction Pathway for Ethanol Steam Reforming on a Ni/SiO2 Catalyst Including Coke Formation. Int. J. Hydrogen Energy 2014, 39, 18820–18834.
  29. Xu, W.; Liu, Z.; Johnston-Peck, A.C.; Senanayake, S.D.; Zhou, G.; Stacchiola, D.; Stach, E.A.; Rodriguez, J.A. Steam Reforming of Ethanol on Ni/CeO2: Reaction Pathway and Interaction between Ni and the CeO2 Support. ACS Catal. 2013, 3, 975–984.
  30. 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.
  31. Grzybek, G.; Greluk, M.; Tarach, K.; Pyra, K.; Słowik, G.; Rotko, M.; Góra-Marek, K. Bioethanol Steam Reforming over Cobalt-Containing USY and ZSM-5 Commercial Zeolite Catalysts. Front. Mater. 2020, 7, 597528.
  32. Sohn, H.; Ozkan, U.S. Cobalt-Based Catalysts for Ethanol Steam Reforming: An Overview. Energy Fuels 2016, 30, 5309–5322.
  33. Erdohelyi, A.; Raskó, J.; Kecskés, T.; Tóth, M.; Dömök, M.; Baán, K. Hydrogen Formation in Ethanol Reforming on Supported Noble Metal Catalysts. Catal. Today 2006, 116, 367–376.
  34. Frusteri, F.; Freni, S.; Spadaro, L.; Chiodo, V.; Bonura, G.; Donato, S.; Cavallaro, S. H2 Production for MC Fuel Cell by Steam Reforming of Ethanol over MgO Supported Pd, Rh, Ni and Co Catalysts. Catal. Commun. 2004, 5, 611–615.
  35. Dalena, F.; Giglio, E.; Marino, A.; Aloise, A.; Giorgianni, G.; Migliori, M.; Giordano, G. Steam Reforming of Bioethanol Using Metallic Catalysts on Zeolitic Supports: An Overview. Catalysts 2022, 12, 617.
  36. Montero, C.; Remiro, A.; Valle, B.; Oar-Arteta, L.; Bilbao, J.; Gayubo, A.G. Origin and Nature of Coke in Ethanol Steam Reforming and Its Role in Deactivation of Ni/La2O3-αAl2O3 Catalyst. Ind. Eng. Chem. Res. 2019, 58, 14736–14751.
  37. Ni, M.; Leung, D.Y.C.; Leung, M.K.H. A Review on Reforming Bio-Ethanol for Hydrogen Production. Int. J. Hydrogen Energy 2007, 32, 3238–3247.
  38. Song, H.; Zhang, L.; Watson, R.B.; Braden, D.; Ozkan, U.S. Investigation of Bio-Ethanol Steam Reforming over Cobalt-Based Catalysts. Catal. Today 2007, 129, 346–354.
  39. Contreras, J.L.; Salmones, J.; Colín-Luna, J.A.; Nuño, L.; Quintana, B.; Córdova, I.; Zeifert, B.; Tapia, C.; Fuentes, G.A. Catalysts for H2 Production Using the Ethanol Steam Reforming (a Review). Int. J. Hydrogen Energy 2014, 39, 18835–18853.
  40. Salge, J.R.; Deluga, G.A.; Schmidt, L.D. Catalytic Partial Oxidation of Ethanol over Noble Metal Catalysts. J. Catal. 2005, 235, 69–78.
  41. Vita, A.; Pino, L.; Italiano, C.; Palella, A. Steam Reforming, Partial Oxidation, and Autothermal Reforming of Ethanol for Hydrogen Production in Conventional Reactors. In Ethanol: Science and Engineering; Elsevier: Amsterdam, The Netherlands, 2018; pp. 159–191. ISBN 9780128114582.
  42. Youn, M.H.; Seo, J.G.; Lee, H.; Bang, Y.; Chung, J.S.; Song, I.K. Hydrogen Production by Auto-Thermal Reforming of Ethanol over Nickel Catalysts Supported on Metal Oxides: Effect of Support Acidity. Appl. Catal. B 2010, 98, 57–64.
  43. Gutierrez, A.; Karinen, R.; Airaksinen, S.; Kaila, R.; Krause, A.O.I. Autothermal Reforming of Ethanol on Noble Metal Catalysts. Int. J. Hydrogen Energy 2011, 36, 8967–8977.
  44. Chen, W.H.; Biswas, P.P.; Ong, H.C.; Hoang, A.T.; Nguyen, T.B.; Dong, C. Di A Critical and Systematic Review of Sustainable Hydrogen Production from Ethanol/Bioethanol: Steam Reforming, Partial Oxidation, and Autothermal Reforming. Fuel 2023, 333, 126526.
  45. Chen, W.H.; Biswas, P.P.; Ubando, A.T.; Park, Y.K.; Ashokkumar, V.; Chang, J.S. Design of Experiment for Hydrogen Production from Ethanol Reforming: A State-of-the-Art Review. Fuel 2023, 342, 127871.
  46. Song, H.; Luo, S.; Huang, H.; Deng, B.; Ye, J. Solar-Driven Hydrogen Production: Recent Advances, Challenges, and Future Perspectives. ACS Energy Lett. 2022, 7, 1043–1065.
  47. Gupta, A.; Likozar, B.; Jana, R.; Chanu, W.C.; Singh, M.K. A Review of Hydrogen Production Processes by Photocatalytic Water Splitting—From Atomistic Catalysis Design to Optimal Reactor Engineering. Int. J. Hydrogen Energy 2022, 47, 33282–33307.
  48. Yao, Y.; Gao, X.; Li, Z.; Meng, X. Photocatalytic Reforming for Hydrogen Evolution: A Review. Catalysts 2020, 10, 335.
  49. Christoforidis, K.C.; Fornasiero, P. Photocatalytic Hydrogen Production: A Rift into the Future Energy Supply. ChemCatChem 2017, 9, 1523–1544.
  50. Kumaravel, V.; Mathew, S.; Bartlett, J.; Pillai, S.C. Photocatalytic Hydrogen Production Using Metal Doped TiO2: A Review of Recent Advances. Appl. Catal. B 2019, 244, 1021–1064.
  51. Bowker, M. Sustainable Hydrogen Production by the Application of Ambient Temperature Photocatalysis. Green. Chem. 2011, 13, 2235–2246.
  52. Escobedo, S.; de Lasa, H. Synthesis and Performance of Photocatalysts for Photocatalytic Hydrogen Production: Future Perspectives. Catalysts 2021, 11, 1505.
  53. Karimi Estahbanati, M.R.; Feilizadeh, M.; Attar, F.; Iliuta, M.C. Current Developments and Future Trends in Photocatalytic Glycerol Valorization: Photocatalyst Development. Ind. Eng. Chem. Res. 2020, 59, 22330–22352.
  54. Rueda-Navarro, C.M.; Ferrer, B.; Baldoví, H.G.; Navalón, S. Photocatalytic Hydrogen Production from Glycerol Aqueous Solutions as Sustainable Feedstocks Using Zr-Based UiO-66 Materials under Simulated Sunlight Irradiation. Nanomaterials 2022, 12, 3808.
  55. Augustin, A.; Chuaicham, C.; Shanmugam, M.; Vellaichamy, B.; Rajendran, S.; Hoang, T.K.A.; Sasaki, K.; Sekar, K. Recent Development of Organic-Inorganic Hybrid Photocatalysts for Biomass Conversion into Hydrogen Production. Nanoscale Adv. 2022, 4, 2561–2582.
  56. Shimura, K.; Yoshida, H. Heterogeneous Photocatalytic Hydrogen Production from Water and Biomass Derivatives. Energy Environ. Sci. 2011, 4, 2467–2481.
  57. Patsoura, A.; Kondarides, D.I.; Verykios, X.E. Photocatalytic Degradation of Organic Pollutants with Simultaneous Production of Hydrogen. Catal. Today 2007, 124, 94–102.
  58. Karimi Estahbanati, M.R.; Babin, A.; Feilizadeh, M.; Nayernia, Z.; Mahinpey, N.; Iliuta, M.C. Photocatalytic Conversion of Alcohols to Hydrogen and Carbon-Containing Products: A Cleaner Alcohol Valorization Approach. J. Clean. Prod. 2021, 318, 128546.
  59. Sola, A.C.; Homs, N.; Ramírez de la Piscina, P. Photocatalytic H2 Production from Ethanol (Aq) Solutions: The Effect of Intermediate Products. Int. J. Hydrogen Energy 2016, 41, 19629–19636.
  60. Barba-Nieto, I.; Caudillo-Flores, U.; Gómez-Cerezo, M.N.; Kubacka, A.; Fernández-García, M. Boosting Pt/TiO2 Hydrogen Photoproduction through Zr Doping of the Anatase Structure: A Spectroscopic and Mechanistic Study. Chem. Eng. J. 2020, 398, 125665.
  61. Schneider, J.; Matsuoka, M.; Takeuchi, M.; Zhang, J.; Horiuchi, Y.; Anpo, M.; Bahnemann, D.W. Understanding TiO2 photocatalysis: Mechanisms and Materials. Chem. Rev. 2014, 114, 9919–9986.
  62. Yang, Y.Z.; Chang, C.H.; Idriss, H. Photo-Catalytic Production of Hydrogen Form Ethanol over M/TiO2 Catalysts (M = Pd, Pt or Rh). Appl. Catal. B 2006, 67, 217–222.
  63. Rusinque, B.; Escobedo, S.; de Lasa, H. Photocatalytic Hydrogen Production under Near-UV Using Pd-Doped Mesoporous TiO2 and Ethanol as Organic Scavenger. Catalysts 2019, 9, 33.
  64. Serafin, J.; Ouzzine, M.; Sreńscek-Nazzal, J.; Llorca, J. Photocatalytic Hydrogen Production from Alcohol Aqueous Solutions over TiO2-Activated Carbon Composites Decorated with Au and Pt. J. Photochem. Photobiol. A Chem. 2022, 425, 113726.
  65. Ibrahim, N.S.; Leaw, W.L.; Mohamad, D.; Alias, S.H.; Nur, H. A Critical Review of Metal-Doped TiO2 and Its Structure–Physical Properties–Photocatalytic Activity Relationship in Hydrogen Production. Int. J. Hydrogen Energy 2020, 45, 28553–28565.
  66. Vitiello, G.; Clarizia, L.; Abdelraheem, W.; Esposito, S.; Bonelli, B.; Ditaranto, N.; Vergara, A.; Nadagouda, M.; Dionysiou, D.D.; Andreozzi, R.; et al. Near UV-Irradiation of CuOx-Impregnated TiO2 Providing Active Species for H2 Production Through Methanol Photoreforming. ChemCatChem 2019, 11, 4314–4326.
  67. Clarizia, L.; Russo, D.; Di Somma, I.; Andreozzi, R.; Marotta, R. Hydrogen Generation through Solar Photocatalytic Processes: A Review of the Configuration and the Properties of Effective Metal-Based Semiconductor Nanomaterials. Energies 2017, 10, 1624.
  68. Karimi Estahbanati, M.R.; Mahinpey, N.; Feilizadeh, M.; Attar, F.; Iliuta, M.C. Kinetic Study of the Effects of PH on the Photocatalytic Hydrogen Production from Alcohols. Int. J. Hydrogen Energy 2019, 44, 32030–32041.
  69. Shiva Kumar, S.; Himabindu, V. Hydrogen Production by PEM Water Electrolysis—A Review. Mater. Sci. Energy Technol. 2019, 2, 442–454.
  70. Coutanceau, C.; Baranton, S. Electrochemical Conversion of Alcohols for Hydrogen Production: A Short Overview. Wiley Interdiscip. Rev. Energy Environ. 2016, 5, 388–400.
  71. Lamy, C.; Jaubert, T.; Baranton, S.; Coutanceau, C. Clean Hydrogen Generation through the Electrocatalytic Oxidation of Ethanol in a Proton Exchange Membrane Electrolysis Cell (PEMEC): Effect of the Nature and Structure of the Catalytic Anode. J. Power Sources 2014, 245, 927–936.
  72. Pethaiah, S.S.; Sadasivuni, K.K.; Jayakumar, A.; Ponnamma, D.; Tiwary, C.S.; Sasikumar, G. Methanol Electrolysis for Hydrogen Production Using Polymer Electrolyte Membrane: A Mini-Review. Energies 2020, 13, 5879.
  73. Caravaca, A.; Sapountzi, F.M.; De Lucas-Consuegra, A.; Molina-Mora, C.; Dorado, F.; Valverde, J.L. Electrochemical Reforming of Ethanol-Water Solutions for Pure H 2 Production in a PEM Electrolysis Cell. Int. J. Hydrogen Energy 2012, 37, 9504–9513.
  74. Yu, J.; González-Cobos, J.; Dappozze, F.; Grimaldos-Osorio, N.; Vernoux, P.; Caravaca, A.; Guillard, C. First PEM Photoelectrolyser for the Simultaneous Selective Glycerol Valorization into Value-Added Chemicals and Hydrogen Generation. Appl. Catal. B 2023, 327, 122465.
  75. De Lucas-Consuegra, A.; Calcerrada, A.B.; De La Osa, A.R.; Valverde, J.L. Electrochemical Reforming of Ethylene Glycol. Influence of the Operation Parameters, Simulation and Its Optimization. Fuel Process. Technol. 2014, 127, 13–19.
  76. Liu, Y.P.; Zhao, S.F.; Guo, S.X.; Bond, A.M.; Zhang, J.; Zhu, G.; Hill, C.L.; Geletii, Y.V. Electrooxidation of Ethanol and Methanol Using the Molecular Catalyst 10−. J. Am. Chem. Soc. 2016, 138, 2617–2628.
  77. Asiri, H.A.; Anderson, A.B. Mechanisms for Ethanol Electrooxidation on Pt(111) and Adsorption Bond Strengths Defining an Ideal Catalyst. J. Electrochem. Soc. 2015, 162, F115–F122.
  78. Lima, F.H.B.; Gonzalez, E.R. Ethanol Electro-Oxidation on Carbon-Supported Pt-Ru, Pt-Rh and Pt-Ru-Rh Nanoparticles. Electrochim. Acta 2008, 53, 2963–2971.
  79. Zhang, B.W.; Sheng, T.; Wang, Y.X.; Qu, X.M.; Zhang, J.M.; Zhang, Z.C.; Liao, H.G.; Zhu, F.C.; Dou, S.X.; Jiang, Y.X.; et al. Platinum-Cobalt Bimetallic Nanoparticles with Pt Skin for Electro-Oxidation of Ethanol. ACS Catal. 2017, 7, 892–895.
  80. Rizo, R.; Sebastián, D.; Lázaro, M.J.; Pastor, E. On the Design of Pt-Sn Efficient Catalyst for Carbon Monoxide and Ethanol Oxidation in Acid and Alkaline Media. Appl. Catal. B 2017, 200, 246–254.
  81. Silva, J.C.M.; De Souza, R.F.B.; Parreira, L.S.; Neto, E.T.; Calegaro, M.L.; Santos, M.C. Ethanol Oxidation Reactions Using SnO2@Pt/C as an Electrocatalyst. Appl. Catal. B 2010, 99, 265–271.
  82. Rizo, R.; Bergmann, A.; Timoshenko, J.; Scholten, F.; Rettenmaier, C.; Jeon, H.S.; Chen, Y.T.; Yoon, A.; Bagger, A.; Rossmeisl, J.; et al. Pt-Sn-Co Nanocubes as Highly Active Catalysts for Ethanol Electro-Oxidation. J. Catal. 2021, 393, 247–258.
  83. Lamy, C.; Rousseau, S.; Belgsir, E.M.; Coutanceau, C.; Léger, J.M. Recent Progress in the Direct Ethanol Fuel Cell: Development of New Platinum-Tin Electrocatalysts. Electrochim. Acta 2004, 49, 3901–3908.
  84. Yang, G.; Zhang, Q.; Yu, H.; Peng, F. Platinum-Based Ternary Catalysts for the Electrooxidation of Ethanol. Particuology 2021, 58, 169–186.
  85. Kowal, A.; Li, M.; Shao, M.; Sasaki, K.; Vukmirovic, M.B.; Zhang, J.; Marinkovic, N.S.; Liu, P.; Frenkel, A.I.; Adzic, R.R. Ternary Pt/Rh/SnO2 Electrocatalysts for Oxidizing Ethanol to CO2. Nat. Mater. 2009, 8, 325–330.
  86. Rodríguez-Gómez, A.; Dorado, F.; Sánchez, P.; de la Osa, A.R. Boosting Hydrogen and Chemicals Production through Ethanol Electro-Reforming on Pt-Transition Metal Anodes. J. Energy Chem. 2022, 70, 394–406.
  87. Serrano-Jiménez, J.; de la Osa, A.R.; Rodríguez-Gómez, A.; Sánchez, P.; Romero, A.; de Lucas-Consuegra, A. Electro-Reforming of Bioethanol Produced by Sugar Fermentation on a Pt-Ni Anodic Catalyst Supported on Graphene Nanoplatelets. J. Environ. Chem. Eng. 2023, 11, 109703.
  88. Khamhaeng, P.; Laosiripojana, N.; Assabumrungrat, S.; Kim-Lohsoontorn, P. Techno-Economic Analysis of Hydrogen Production from Dehydrogenation and Steam Reforming of Ethanol for Carbon Dioxide Conversion to Methanol. Int. J. Hydrogen Energy 2021, 46, 30891–30902.
  89. Khamhaeng, P.; Kim-Lohsoontorn, P. Performance and Cost Analysis of Hydrogen Production from Steam Reforming and Dehydrogenation of Ethanol. In IOP Conference Series: Materials Science and Engineering; IOP Publishing Ltd.: Bristol, UK, 2020; Volume 991.
  90. Tóth, M.; Varga, E.; Oszkó, A.; Baán, K.; Kiss, J.; Erdohelyi, A. Partial Oxidation of Ethanol on Supported Rh Catalysts: Effect of the Oxide Support. J. Mol. Catal. A Chem. 2016, 411, 377–387.
  91. Chiu, W.-C.; Horng, R.-F.; Chou, H.-M. Hydrogen Production from an Ethanol Reformer with Energy Saving Approaches over Various Catalysts. Int. J. Hydrogen Energy 2013, 38, 2760–2769.
  92. Hung, C.C.; Chen, S.L.; Liao, Y.K.; Chen, C.H.; Wang, J.H. Oxidative Steam Reforming of Ethanol for Hydrogen Production on M/Al2O3. Int. J. Hydrogen Energy 2012, 37, 4955–4966.
  93. Silva Júnior, M.E.; Palm, M.O.; Duarte, D.A.; Catapan, R.C. Catalytic Pt/Al2O3 Monolithic Foam for Ethanol Reforming Fabricated by the Competitive Impregnation Method. ACS Omega 2023, 8, 6507–6514.
  94. Jovic, V.; Al-Azri, Z.H.N.; Chen, W.T.; Sun-Waterhouse, D.; Idriss, H.; Waterhouse, G.I.N. Photocatalytic H2 Production from Ethanol-Water Mixtures over Pt/TiO2 and Au/TiO2 Photocatalysts: A Comparative Study. Top. Catal. 2013, 56, 1139–1151.
  95. Al-Azri, Z.H.N.; Jovic, V.; Chen, W.-T.; Sun-Waterhouse, D.; Metson, J.B.; Waterhouse, G.I.N. Performance Evaluation of Pd/TiO2 and Pt/TiO2 for Hydrogen Production from ethanol-Water Mixtures. Int. J. Nanotechnol. 2014, 11, 695–703.
  96. Di Benedetto, A.; Portarapillo, M.; Landi, G.; Luciani, G. Process for Green Hydrogen Production. WO Patent WO2023105545A1, 7 December 2022.
  97. Di Nardo, A.; Portarapillo, M.; Russo, D.; Luciani, G.; Landi, G.; Ruoppolo, G.; Pezzella, A.; Di Benedetto, A. Cyan Hydrogen Process: A New Route for Simultaneous Hydrogen Production and Carbon Valorisation. ACS Omega, 2023; in press.
More
Information
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , , , , , ,
View Times: 129
Revisions: 4 times (View History)
Update Date: 22 Dec 2023
1000/1000