You are on page 1of 16

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/348456784

Progress and Prospects of Hydrogen Production: Opportunities and


Challenges

Article  in  Journal of Electronic Science and Technology · January 2021


DOI: 10.1016/j.jnlest.2021.100080

CITATIONS READS

72 1,212

5 authors, including:

Bing Zhang
Shenyang University of Technology
57 PUBLICATIONS   823 CITATIONS   

SEE PROFILE

All content following this page was uploaded by Bing Zhang on 19 August 2021.

The user has requested enhancement of the downloaded file.


Journal of Electronic Science and Technology 19 (2021) 100080

Contents lists available at ScienceDirect

Journal of Electronic Science and Technology


journal homepage: www.keaipublishing.com/en/journals/journal-of-electronic-
science-and-technology/

Progress and prospects of hydrogen production: Opportunities


and challenges
Bing Zhang a, *, Sui-Xin Zhang a, Rui Yao a, Yong-Hong Wu a, Jie-Shan Qiu b
a
School of Petrochemical Engineering, Shenyang University of Technology, Liaoyang, 111003, China
b
College of Chemical Engineering, Beijing University of Chemical Technology, Beijing, 100029, China

A R T I C L E I N F O A B S T R A C T

Publishing editor: Yu-Lian He This study presents an overview of the current status of hydrogen production in relation to the
global requirement for energy and resources. Subsequently, it symmetrically outlines the advan-
Keywords: tages and disadvantages of various production routes including fossil fuel/biomass conversion,
Electrolysis water electrolysis, microbial fermentation, and photocatalysis (PC), in terms of their technologies,
Fermentation
economy, energy consumption, and costs. Considering the characteristics of hydrogen energy and
Hydrogen production
the current infrastructure issues, it highlights that onsite production is indispensable and conve-
Photocatalysis (PC)
Process integration nient for some special occasions. Finally, it briefly summarizes the current industrialization situ-
Renewable resource ation and presents future development and research directions, such as theoretical research
Sustainable energy strengthening, renewable raw material development, process coupling, and sustainable energy
use.

1. Introduction

Energy is a fundamental requirement for human survival, social civilization, and economic development. Traditional fossil fuels,
such as coal, petroleum, and natural gas, have been used for over 20 decades, resulting in excessive energy consumption, unrestrained
exploitation, and considerable waste. Consequently, these nonrenewable resources are rapidly approaching depletion and exhaustion.
Particularly, the successive expansion of the global population and the rapid economic development have resulted in an increased daily
energy demand, worsening the energy crisis issue [1]. Additionally, the excessive exploitation and overuse of fossil fuels have caused
serious environmental pollution. Hence, most countries are keen to find an alternative clean energy source [2,3].
Hydrogen, as a clean energy carrier for heat and electricity, has many appealing characteristics, including a large storage capacity,
high energy conversion, cleanliness and environmental friendliness, renewable production, vast specific energy, zero emissions, wide
sources, reliability, and easy storage and regeneration [4,5]. Thus, it is considered to be the cleanest and most promising energy source
of the 21st century. Additionally, hydrogen is an essential chemical raw material, widely used in large-scale industrial processes
including ammonia synthesis, petroleum refining, and the water-gas shift reaction. From the relationship between the supply and
demand shown in Fig. 1, it can be inferred that hydrogen is extremely versatile, and it offers promising prospects. Thus, hydrogen
production has received considerable global attention.
Since hydrogen does not exist as a molecule in nature, it is produced by the conversion of certain source materials containing the
hydrogen element (e.g., water or carbohydrate). Presently, over 96% of the global hydrogen is produced from traditional fossil sources,

* Corresponding author.
E-mail addresses: zhangbing@sut.edu.cn (B. Zhang), 18856330512@qq.com (S.-X. Zhang), 1063983065@qq.com (R. Yao), wuyh@sut.edu.cn
(Y.-H. Wu), qiujs@mail.buct.edu.cn (J.-S. Qiu).

https://doi.org/10.1016/j.jnlest.2021.100080
Received 26 June 2020; Received in revised form 12 September 2020; Accepted 31 December 2020
Available online 11 January 2021
1674-862X/© 2021 University of Electronic Science and Technology of China. Publishing Services provided by Elsevier B.V. on behalf of KeAi
Communications Co. Ltd. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
B. Zhang et al. Journal of Electronic Science and Technology 19 (2021) 100080

Fig. 1. Relationship between the supply and demand of hydrogen.

of which steam reforming of natural gas contributes 48%, naphtha reforming 30%, and coal gasification 18% [6]. Unfortunately, most of
the traditional technologies for hydrogen production from fossil fuels are associated with serious environmental pollution and high
energy consumption. Therefore, increasing attention has been paid to the application of new technologies for hydrogen production from
renewable and nuclear sources, with increasingly strict and relevant environmental protection regulations worldwide [7]. Such tech-
nologies include water electrolysis, biomass gasification, and nuclear thermal/chemical routes. Regardless, the hydrogen-production
process must consume a certain amount of materials and energy via either a traditional or new technology, which will inevitably
have an impact on the environment [8]. Only by fully understanding each section of the hydrogen-production process can we objectively
evaluate the economic and environmental benefits of various hydrogen-production technologies.
In the following section, the current hydrogen-production routes are discussed briefly.

Fig. 2. Hydrogen-production routes.

2
B. Zhang et al. Journal of Electronic Science and Technology 19 (2021) 100080

2. Hydrogen-production routes

In the industrial chain of hydrogen energy (i.e., hydrogen production, storage and transportation, hydrogen fueling, and applica-
tions), hydrogen production is the most important segment that impacts the potential applications. Herein, we outline the technologies
reported in the literature, including reforming, gasification, water electrolysis, microbial fermentation, and photocatalysis (PC), as well
as some other conversion methods involving fossil fuels or biomass, which have been summarized in Fig. 2.

2.1. Source materials

There are two major source materials for hydrogen production: Fossil fuels and biomass, which will be discussed detailedly in the
following.

2.1.1. Fossil fuel source


A large variety of fossil fuels, such as natural gas, oil, and coal, can be applied to produce hydrogen via steam reforming or partial
oxidation. The increasing demand for hydrogen, together with the growing coal chemical industry, deterioration of crude oil, upgrading
of fuel oil quality, and progress in hydrogen energy technologies, has resulted in benefits and challenges for small- and large-scale coal-
to-hydrogen production.
In China, where more coal is consumed than oil, coal gasification will remain the main resource for large-scale hydrogen production.
From the viewpoint of chemistry, coal is a complex material with a highly variable composition, which can be converted into a variety of
products. Coal gasification is a method applied to generate power, liquid fuels, chemicals, and hydrogen. Specifically, hydrogen is
produced by reacting coal with oxygen and steam at high pressure and temperatures to form synthesis gas, a mixture primarily
comprised of carbon monoxide and hydrogen. Thereafter, hydrogen is recovered by pressure swing adsorption (PSA), which has the
advantages of wide adaptability of feeding gas, cleanliness, energy saving, a high degree of automation, good reliability and flexibility,
convenient startup and shutdown, high hydrogen purity, etc. The future directions for this technique are as follows: 1) Improving the
performance of the adsorbent, 2) optimizing the adsorption tower structure to strengthen the utilization rate of the adsorbent and
eliminate the influence of pressure and temperatures on the stability, and 3) separating the gas mixture by coupling the low-temperature
absorption and membrane technology [9]. Additionally, it is very important to study high-efficiency catalysts (e.g., iron-based oxides)
for the water-gas shift reaction [10].
Hydrogen production from natural gas occurs mainly via steam reforming, partial oxidation, autothermal reforming, and chemical
chain reforming. The key factors influencing this technology are the catalyst, reactor type, impurities in the raw materials, choice of
heating fuels during the reaction, carbon capture in the product, and choice of the hydrogen purification equipment. Osman investigated
the two main mechanisms of partial oxidation of methane, namely direct partial oxidation and combustion and the reforming reaction,
based on recent kinetic studies [11]. Gao et al. discussed the preparation of catalysts, optimization of operation conditions, and influence
of impurities in the dry reforming of biogas (methane and carbon dioxide), and they provided guidelines for its industrial applications in
hydrogen energy generation [12].
Apart from methane, some other hydrocarbons can be used to produce hydrogen by thermo-catalytic decomposition (TCD).
Muhammad et al. discussed the practice of co-feeding methane with other hydrocarbons, specifically ethylene, propylene, hydrogen
sulfide, and ethanol, with a detailed overview of TCD reaction kinetics over an empirical model based on the power law [13].
Ethanol has also attracted significant attention from researchers and governments for hydrogen production in recent years, because it
can be produced from either fossil or renewable biomass raw materials. Taghizadeh and Aghili recommended the integration of the
catalytic reaction and separation process into a single process in a single membrane reactor to significantly improve the conversion and
hydrogen yields and reduce the reactor volume. They also presented a brief overview of the effects of different catalysts (such as precious
metals, non-precious metals, and bimetallic catalysts) [14]. Ogo and Sekine developed a novel electrocatalytic reaction system involving
conducting a catalytic reaction in an electric field for low-temperature catalytic ethanol steam reforming with low-grade waste heat
[15]. In this electrocatalytic reaction system, hydrogen was produced efficiently from ethanol using a small amount of electricity, even
at low temperatures.

2.1.2. Biomass source


Compared with fossil fuels, biomass is more promising for hydrogen production because of its large reserves and supply, high annual
output, and easy oxidation.
Thus far, hydrogen has been produced from various types of biomass via suitable methods, including PC of municipal solid waste,
agricultural waste, forest residues [16], polyols, carboxylic acids, and saccharides [17]; thermochemical conversion of wood waste,
sawdust, olivine particles, lignin, cellulose, and biomass char [18]; fermentation of microalgae and cassava (see subsection 2.3) [19,20];
steam reforming of gasified biomass tar [21], biomethane (biogas) [22], and bioethanol [23]; supercritical water gasification of
wastewater sludge [24]; aqueous phase reforming of cellulose, sugars, and polyols [25].
However, most of these processes tend to form more complex and undegradable derivatives after hydrogen production. This limits
the application of biomass for hydrogen production. Therefore, the selective conversion of biomass into chemicals or oil is essential for
reducing the production cost of the desired product and achieving the full utilization of raw biomass materials.

3
B. Zhang et al. Journal of Electronic Science and Technology 19 (2021) 100080

2.2. Water electrolysis

Electrolysis is the process of applying an electric current to dissociate water or a water-containing solution into hydrogen gas and
oxygen gas, as illustrated in Fig. 3 [26]. The electric potential shifts positive ions (Hþ) toward the cathode and negative ions (OH)
toward the anode. There are different water electrolysis processes, i.e., alkaline water electrolysis, proton exchange membrane water
electrolysis, solid oxide water electrolysis, and alkaline anion exchange membrane (AEM) water electrolysis [27]. Among them, alkaline
water electrolysis has been successfully established on a pilot demonstration project by the Dalian Institute of Chemical Physics, Chinese
Academy of Sciences, of which the highest energy efficiency that can be reached is ~88%.
Hydrogen production from electrolyzed coal slurry is superior to the traditional electrolysis water process in terms of energy con-
sumption and production efficiency. This technology can simultaneously achieve the purification of the ore during the electrolysis
process, which is worthy of promotion and development. The key future directions are as follows: A reduction of energy consumption in
the electrolysis of coal water slurry using renewable energy; an in-depth study of the reaction mechanism to improve the efficiency of
hydrogen production and achieve conversion via the coupling of chemical energy and electrical energy; an improvement of the electrode
stability and corrosion resistance for improved durability to reduce the electrode cost; an enhancement of the reaction efficiency by
developing new catalytic electrodes and catalysts [28,29].
It is also believed that the utilization of AEM water electrolysis can enable the replacement of conventional noble metal electro-
catalysts (Pt, Pd, Ru, and Ir) with low-cost transition metals. Although it is still a developing technology, AEM electrolysis has attracted
special attention because of its high power efficiency, membrane stability, robustness, handling ease, and the low-cost hydrogen-pro-
duction process [30].
Besides the expensive electrode materials, another bottleneck of hydrogen production from water electrolysis is high energy con-
sumption because of the increase in the electrolysis voltage caused by the bubbles generated during the electrolysis process [31]. The
incorporation of hydrocarbons in water electrolysis can reduce energy consumption. The probable future direction for electrodes should
be inexpensive metals or nonmetal composite materials, such as Ni. Moreover, the method of exhausting gas bubbles should be the focus
of future research.

2.3. Microbial fermentation

Biohydrogen generated from a wide variety of biomass has garnered substantial research interest in several fields. One of the major
goals in this field is to develop technically sound and effective approaches for significantly improving bioconversion and biohydrogen
yields.

Fig. 3. Schematic of water electrolysis [26].

4
B. Zhang et al. Journal of Electronic Science and Technology 19 (2021) 100080

Fermentative conversion of biomass gained significant interest during the last decade. Biohydrogen is not only an efficient trans-
portation fuel but also a low-carbon source of heat and electricity. Microbe-assisted conversion (bioconversion) can occur either in the
presence or absence of light, i.e., photo fermentation or dark fermentation, respectively [32]. The dark-/photo-fermentation technology
is a new approach for increasing the hydrogen yield and energy recovery from organic waste [33].
Fig. 4 gives the two pathways of biohydrogen production [34]. Generally, photo fermentation is associated with a high yield and low
productivity, while dark fermentation is associated with a low yield and high productivity. Moreover, the hydrogen-production effi-
ciency is tightly associated with the selected process factors, including the pretreatment conditions, raw materials, and enzymes or
bacteria.
So far, several fungi species and enzymes have been exploited for hydrogen production from a large variety of biomass or organic
substances. These include microalgae (including cyanobacteria and green algae) [35], [FeFe]-hydrogenases [36], nitrogenase [37],
Escherichia coli [38], cyanobacteria [39], and the genus Clostridium [40]. Among them, biomass is particularly interesting because of its
abundant availability, low costs, mild reaction conditions (usually at ambient temperature and pressure), environmental friendliness,
energy saving, and zero consumption of mineral resources. To date, several types of biomass have been validated in the literature,
including organic waste [41], food waste [42], carbohydrate-rich biomass [43], sugars [44], lignocellulose [45], sewage sludge [46],
macroalgal [47], and crop residues [48].
Moreover, other factors, such as the fermentation mode [32], culture system pretreatment [49], and additives [50], influence the
hydrogen-production efficiency.
Presently, the research trends of this technology are mainly focused on the following points: 1) Breeding bacterial sources with anti-
inhibition and excellent performance to improve the yield and efficiency of hydrogen production; 2) expanding more biomass types as
raw materials, particularly for annual large-volume production (moreover, the simultaneous realization of multiple objectives, such as
environmental protection and high values of biomass resources, by considering the process of hydrogen production from renewable raw
materials, industrial and agricultural wastewater, and waste materials, can be achieved); 3) developing more clean and efficient pro-
duction processes by optimizing the configuration of the fermentative reactor and process. Thus, the economic feasibility of this
technique can be effectively improved.

2.4. PC

Solar water splitting is a promising approach for transforming sunlight into renewable, sustainable, and green hydrogen energy using
photocatalytically active materials [51]. There are three representative methods for transforming solar radiation into molecular
hydrogen: PC, photoelectrochemical (PEC), and photovoltaic-phtotoelectrochemical (PV-PEC) routes (Fig. 5) [52-54].
PEC cells can directly convert solar energy to hydrogen fuels by water photoelectrolysis, where the functions of both light-harvesting
and electrolysis are combined. In such devices, photocathodes and photoanodes carry out the hydrogen evolution reaction (HER) and
the oxygen evolution reaction (OER), respectively. The focus of such a device is on photocathodes for HER, which are traditionally metal
oxides, III-V group and II-VI group semiconductors, silicon, and copper-based chalcogenides as photoactive materials [55]. Fereidooni
et al. demonstrated the viability of hydrogen production through electrolysis, supported by a photovoltaic power system for economic
feasibility, with an annual output of 20 kW (the photovoltaic power station is located in Yazd City in Iran) by both experimental studies
and simulations [56]. Ahmed and Dincer discussed the possible commercialization of viable PEC reactors [57]. Varas-Concha et al. [58]
analyzed the effects of five factors on photocatalytic hydrogen production by photoreforming organic compounds. The analyzed factors
were the presence of Au as a cocatalyst, type of alcohol applied as the electron donor, intensity of ultraviolet (UV) light, electron donor
concentration, and nanoparticle concentration. A main and interaction effect analysis is presented with reduced fixed-effect models for

Fig. 4. Two microbial pathways for biohydrogen production [34].

5
B. Zhang et al. Journal of Electronic Science and Technology 19 (2021) 100080

Fig. 5. Illustrations for hydrogen production via solar water splitting: (a) PC, (b) PEC, and (c) PV-PEC [54].

three responses: Total hydrogen generation, catalyst productivity, and electron donor productivity. The presence of Au as a cocatalyst,
the intensity of UV light, and their interaction had the highest effect. The best configuration allowed us to achieve the catalyst pro-
ductivity of 2925 μmol(H2) g1h1.
Regardless, the most important factor for realizing hydrogen production is the photocatalyst. The current catalysts used for the
photolysis of water to produce hydrogen mainly include metal oxides, metal sulfides, metal nitrogen (nitrogen oxide) compounds,
graphite carbonitrides, and new heterostructure photocatalysts [59-62]. Most of them have a low utilization rate of solar energy and low
recombination of photogenerated electron-hole pairs. Moreover, another vital limitation of photocorrosion or cycling stability must be
tackled before practical applications [63]. These problems lead to low hydrogen-production efficiency, which severely limits the actual
photocatalyst use. Therefore, the development of inexpensive, environmentally friendly, and high-performance photocatalysts with a
relatively long wavelength spectral response is essential for photocatalytic hydrogen production [64].
TiO2, as a typical metal oxide, has been extensively studied for water photolysis to produce hydrogen. Additionally, metal sulfides,
such as CdS, ZnS, and their solid solutions, show good catalytic activity in hydrogen production by water photolysis, because of their
narrow and suitable bandgaps. Metal nitrogens (nitrogen oxides) have an ideal visible-light water-splitting energy band structure, which
requires modification to reflect the water-splitting activity. Graphite carbonitrides, a new type of nonmetallic visible-light catalyst,
present stupendous potential in hydrogen production. Recently, carbon nanomaterial-modified photocatalysts (e.g., fullerenes, carbon
nanotubes, graphene, carbon nanofibers, and carbon quantum dots) have gained significant interest because of their peculiarly
controllable morphology [65]. Moreover, the use of earth-abundant and inexpensive co-catalysts as replacements of noble metals is
necessary for photocatalytic hydrogen production to overcome the main challenges of the high photocatalyst cost and low efficiency for
photocatalytic hydrogen production [66].
Different from inorganic semiconductor photocatalysts, organic semiconductor materials have the advantages of diverse synthesis
methods, easy functional modification, facile adjustment of energy band and electronic structure, etc. [67]. These properties enable
their applications in the field of photocatalytic hydrogen production [68]. Particularly, the organic conjugated microporous polymer
materials developed in recent years have both the semiconductor characteristics of traditional conjugated polymers and the porous
characteristics of high specific-surface-area materials [69].
Apart from the frequently used water, some other resources can be adopted as raw materials for photocatalytic hydrogen production,
such as liquid waste [70], glycerol [71], and many other types of biomass.

2.5. Other storage or conversion methods

In practice, the transportation and storage of hydrogen are difficult because of its hard compression and liquefaction, as well as its
flammable and explosive properties. This restricts its application range. The use of hydrogen-containing compounds (hydrocarbons,
alcohols, ethers, etc.) for in situ online reforming for hydrogen production is considered the most realistic solution. As discussed, there
are many reforming technologies for hydrogen production, such as methane, methanol, ethanol, and dimethyl ether (DME) reforming
methods. Among them, the temperature required for the reforming reaction and energy consumption of methane and ethanol to produce
hydrogen is relatively high. Additionally, methanol is toxic and detrimental to the environment and human beings. Moreover, methanol
is a liquid; thus, it requires additional gasification equipment, which increases the operating cost. Differently, DME is nontoxic, non-
corrosive, environmentally friendly, and easy to store and transport.
Recently, some other species, including acetic acid, formic acid, hydrous hydrazine, ammonia borane, ammonia, and sulfuric acid,
were proposed as good candidates for the applications in hydrogen conversion.
Acetic acid is the major component of bio-oil derived from biomass and has been extensively studied for hydrogen generation via
catalytic steam reforming [72]. The catalyst selection is one of the most important processes in this technology. The heterogeneous
catalysts include the supported and unsupported types. The supported catalysts are far superior to the unsupported catalysts in terms of
catalytic performance for hydrogen production.
Formic acid, the simplest carboxylic acid, has emerged as one of the most promising candidates for the application as liquid organic

6
B. Zhang et al. Journal of Electronic Science and Technology 19 (2021) 100080

hydrogen carriers. Similarly, the preparation of high-efficiency catalysts is the most concerning hurdle in the conversion of hydrogen
from formic acid [73-75].
Subsequently, the desired hydrogen can be fed to a fuel cell, again producing electricity. There are many methods for improving the
efficiency of this process, particularly for the combination of the electrolytic process with other non-electrochemical processes. One of
them is the thermochemical hybrid sulfur cycle (also known as the Westinghouse cycle). This cycle combines a thermochemical step
(H2SO4 decomposition) with an electrochemical one, where hydrogen is produced from the oxidation of SO2 and H2O (SO2 depolari-
zation electrolysis, conducted at a considerably lower cell voltage compared with that of conventional electrolysis) [76].
Hydrous hydrazine (N2H4⋅H2O) is a promising hydrogen storage material for fuel cells because of its high hydrogen content and easy
transportation and storage; furthermore, N2 is the only byproduct of its complete decomposition. Over the past two decades, in response
to the global clean and sustainable energy strategies, nanocatalysts have been intensively investigated for the decomposition of hydrous
hydrazine with high selectivity for hydrogen production [77].
Ammonia borane is another satisfactory hydrogen storage material with the characteristics of high hydrogen storage density (152.9
g/L), mild hydrogen release conditions, nontoxicity, stable storage at room temperature, and easy storage and transportation. The
release of hydrogen from ammonia borane can be fulfilled by three protocols, i.e., pyrolysis, alcoholysis, or hydrolysis. For the pyrolysis
of ammonia borane, attention is paid to reducing the temperature and inhibiting the formation of gaseous byproducts. The research hot
spot of hydrogen production by hydrolysis or alcoholysis is the binary or ternary non-noble metal nanocore shell or supported catalysts.
Compared with the pyrolysis of ammonia borane, hydrolysis or alcoholysis is more practical because of the mild conditions and fast
hydrogen release. The largest challenge associated with ammonia borane as a hydrogen storage material is its regeneration, because the
derived byproducts after its decomposition cannot be directly hydrogenated again [78].
The first step of the ammonia decomposition route is the evaporation of liquid ammonia to ammonia gas through a preheater. This is
followed by decomposition into the mixed gas containing 75% hydrogen and 25% nitrogen through a decomposition furnace packed
with a catalyst at a certain temperature. The catalysts mainly include supported ruthenium, nickel, and iron, as well as carbides and
nitrides. Ruthenium-based catalysts exhibit high catalytic activity only under the condition of high loading and high alkalinity on the
catalyst surface. Consequently, more precious metal ruthenium and an anticorrosion reactor are required, which largely elevates the
production cost. Researchers have explored some inexpensive catalysts, such as iron, nitride, and nickel. It has been found that nickel-
based catalysts have the advantages of low costs, easy preparation, good stability, and broad application prospects, although nickel
exhibits slightly less catalytic activity than ruthenium [79]. The research direction of this technology is on designing new catalysts for
ammonia decomposition at low pressure, low temperatures, and high activity to reduce energy consumption.
Nuclear hydrogen production is one of the most prospective methods for the efficient production of CO2-free hydrogen on a large
scale with water electrolysis or by thermochemical processes [80]. In today’s light-water reactors, the total process efficiency of con-
verting nuclear heat into hydrogen is about 25%, while the total efficiency of power generation by reactors used for water electrolysis is
about 33%. High-temperature reactors can produce hydrogen with up to 50% efficiency using the combination of steam electrolysis and
thermochemical or hybrid processes [81,82]. One typical iodine-sulfur (IS) process includes the Bunsen reaction and the characteristic
separation of an HI/I2/H2SO4/H2O system, the purification of HIx and sulfuric acid phases, the electro-electrodialysis stacks for HI acid
preconcentration, and the catalysts used for HI and SO3 decomposition [83]. Fig. 6 presents a simple schematic of the IS cycle [84]. The
current operating light-water or heavy-water reactors and gas-cooled reactors have a maximum temperature of 350  C. Thus, these
reactors can be used for low-temperature electrolysis. Other than the economic issues, the nuclear hydrogen technology does not have
any major issues for hydrogen production.
Chemical looping reforming is also applicable for direct hydrogen production in the presence of solid metal oxides as oxygen carriers
to replace the water vapor or pure oxygen required in traditional reforming processes. As exemplified in Fig. 7 [85], this method can
directly convert the fuel into high-purity synthesis gas or carbon dioxide and water. Furthermore, the reduced metal oxides can be
regenerated with the water vapor and directly generate hydrogen [86]. The whole process realizes near-zero energy consumption for the
in situ hydrogen separation. According to the different products and heating methods, the direct hydrogen-production system of

Fig. 6. Scheme of the sulfur-iodine cycle [84].

7
B. Zhang et al. Journal of Electronic Science and Technology 19 (2021) 100080

Fig. 7. Scheme of the chemical looping reforming of methane [85].

chemical looping reforming can be categorized into two types: Double-bed system and three-bed system [85]. The key factors for this
method are the selection of oxygen carriers and the reactors to ensure the appropriate gas-solid contact mode, according to the char-
acteristics of different raw materials and products.

3. Current status of industrialization

According to the public reports, the total global demand for hydrogen has rapidly increased from 255.3 billion cubic meters in 2013
to 324.8 billion cubic meters in 2020, an increase of 27.2% [87]. From the previous analysis, it is clear that hydrogen, as a secondary
energy source, should be extracted from substances, such as hydrocarbons and water, through an energy conversion process, such as
thermochemistry and water electrolysis. Among them, thermochemical processes using fossil as raw materials are particularly used in
many industries, mainly including steam reforming of hydrocarbons, partial oxidation of heavy oil, coal gasification, and water
electrolysis.
Unlike laboratory research, industrial applications should comprehensively consider various factors, including the source of raw
materials, technical environmental protection, investment, and operation costs [88]. Presently, the representative processes are natural
gas steam reforming, coal gasification, and water electrolysis. Basically, the production cost of natural gas steam conversion is lower
than that of water electrolysis for hydrogen production under the conditions of different scales, particularly for large-scale plants. When
the scale of hydrogen production is small, the investment required for water electrolysis is lower than that for natural gas steam
reforming. When the scale of hydrogen production is large, the conversion of natural gas steam reforming is lower than that of water
electrolysis. Therefore, it is appropriate to use water electrolysis to produce hydrogen when the raw material supply is not limited, and
the demand for hydrogen is low. Oppositely, it is better to use natural gas steam reforming to produce hydrogen when the demand for
hydrogen is large. Additionally, considering the hydrogen-production cost only, coal gasification is superior to natural gas steam
reforming [89].
Currently, hydrocarbon steam reforming is the most used method globally, utilized in more than 90% of industrial hydrogen-
production plants. The technology was originally invented by Badische Anilin-und-Soda-Fabrik (BASF) in 1926, and Imperial Chemi-
cal Industries (ICI) first realized industrialization in the 1930s [90-92]. The raw materials include natural gas, refinery gas, naphtha,
liquefied gas, and various kinds of hydrocarbon-rich gas. In a reaction system, multiple parallel reactions and series reactions occur
simultaneously. Hydrogen is produced through detoxification purification, water steam conversion, high-temperature shift reactions,
low-temperature shift reactions, decarbonization, and methanation of the raw materials. Because the process is rather mature, the
hydrogen yield over a unit of raw material consumption is relatively high. Nevertheless, the process route is too long, and the purity of
the generated hydrogen is only ~95%.
Heavy oil is the residue produced in petroleum refining, which mainly includes atmospheric residues, vacuum residues, and fuel oil
produced by petroleum deep processing. Hydrogen is generated by the oxidation of heavy oil with oxygen and steam at certain pressure.
Compared with methane, heavy oil has a higher hydrocarbon-to-hydrogen ratio. Therefore, the hydrogen produced by the partial
oxidation of heavy oil mainly originates from steam and carbon monoxide, of which steam contributes 69% [93]. Different from natural
gas steam reforming, the partial oxidation of heavy oil requires additional air separation equipment to produce pure oxygen.
Coal gasification is conducted via the reaction of coal or coke with steam, air, or oxygen at a high temperature to yield a gas mixture
of hydrogen, carbon monoxide, and carbon dioxide. After cooling, dust removal, and desulfurization, the gas mixture undergoes a shift
reaction with steam. Thereafter, a high percentage of the carbon monoxide content is converted into hydrogen and carbon dioxide.
Finally, highly purified hydrogen is recovered by PSA [94]. The representative coal gasification technologies include the dry coal
powder gasification process of Royal Dutch Shell, the coal water slurry gasification process of American General Electric Company and
the GSP (a trademark owned by Siemens Ltd.) gasification technology. All these companies utilize the process of pressurized pure
oxygen fluidized bed liquid slag gasification.

8
B. Zhang et al. Journal of Electronic Science and Technology 19 (2021) 100080

Water electrolysis has been applied industrially for nearly 120 years since 1900. When an aqueous solution of an inorganic acid or
alkali metal hydroxide conducts a direct current using two electrodes that do not react chemically, the cathode generates hydrogen gas,
and the anode generates oxygen gas. Representative companies are German Lurgi Company, Canada Toronto Electrolyzer Co., Ltd., and
Norway Hydro Company. The efficiency of hydrogen production by water electrolysis is 75%–85%. The process is simple and envi-
ronmentally benign; however, the energy consumption is significant, approximately 4.5 kWh⋅m3 to 5.5 kWh⋅m3. The electricity
expenditure accounts for ~80% of the hydrogen-production cost [95].
Hydrogen production from methanol conversion includes methanol decomposition and steam reforming. It involves the dehydro-
genation of methanol at a low temperature to produce methyl formate, formic acid, and hydrogen. Furthermore, the intermediates,
methyl formate and formic acid, decompose to carbon monoxide and hydrogen at high temperatures [96]. The commonly used catalysts
for the reaction include Ni-based catalysts, Cu-based catalysts, and supported precious metal catalysts. Ni-based catalysts have good
stability, a wide application range, and good anti-poison properties; however, they are associated with low activity at low temperatures,
poor selectivity, and high contents of carbon monoxide and methane in the product. Cu-based catalysts have good catalytic activity and
selectivity, although they are easily poisoned. For supported precious metal catalysts, alumina is mostly used as a carrier, platinum and
palladium as active components, and rare earth metals cerium and lanthanum as modifiers; furthermore, they have the advantages of
excellent catalytic activity, selectivity, and stability, and high resistance to poisons and heat, although they are expensive [97]. Typi-
cally, such technologies have been supplied by Haldor Topsoe, Air Liquide, Catalysts and Chemicals Europe, IHI Corporation, Compa~ nía
Espa~nola de Sistemas Aeronauticos (CESA), and Technipetrol.

4. Research and development directions

4.1. Comprehensive evaluation

Besides the rapid increase in demand for hydrogen, the market shows a strong trend toward the diversification of the hydrogen-
production scale. This drives the development of new hydrogen-production technologies, such as photocatalytic water splitting, solar
energy water splitting, nuclear electrolysis water, biomass conversion, hydrogen sulfide decomposition, photochemical cells, aluminum-
alkali redox reactions, phenol steam conversion, and sodium borohydride hydrolysis. Although these technologies are not advanced
enough to replace the present industrial hydrogen-production technologies, they have significant scope for the improvement from the
perspective of long-term development. Future research should focus on achieving green, efficient, convenient, safe, and low-cost
hydrogen-production technologies [98,99].
Overall, microbial fermentation is associated with low efficiency, PC with harsh conditions to initiate the reaction, and water
electrolysis with high energy consumption and costs [100,101]. Table 1 summarizes the capital costs and hydrogen costs of various
methods, where the data are adapted from Ref. [82].
Each technology needs to be comprehensively evaluated according to some general indicators, such as the cleanness index and life
cycle cost, considering the technique, economy, and environment aspects [102,103]. It is particularly important to do an objective
comparative study of various technologies to provide a technical guide for the healthy development of the hydrogen-production
industry.
Xie et al. [104] discussed the life cycle assessment of traditional hydrogen-production technologies (coal gasification, natural gas,
etc.) and new hydrogen-production technologies (thermochemistry, renewable energy power generation, biomass gasification, etc.)
based on the situation in China. As shown in Table 2 with data being adapted from Ref. [104] based on unit hydrogen production, the
wind power technology is the most environmentally friendly method for minimal equivalent greenhouse gas emission, and nuclear
thermal chemistry has the potential for large-scale applications in the future for the lowest energy consumption.

Table 1
Comparison of the capital costs and hydrogen costs of various hydrogen-production methods.
Process Energy source Feedstock Capital cost (M$) Hydrogen cost ($/kg)

Steam methane reforming Fossil fuels Natural gas 180.7 to 226.4 2.08 to 2.27
Coal gasification Fossil fuels Coal 435.9 to 545.6 1.34 to 1.63
Autothermal reforming of methane Fossil fuels Natural gas 183.8 1.48
Methane pyrolysis Internally generated steam Natural gas – 1.59 to 1.70
Biomass pyrolysis Internally generated steam Woody biomass 3.1 to 53.4 1.25 to 2.20
Biomass gasification Internally generated steam Woody biomass 6.4 to 149.3 1.77 to 2.05
Direct bio-photolysis Solar Water þ algae 50 $/m2 2.13
Indirect bio-photolysis Solar Water þ algae 135 $/m2 1.42
Dark fermentation – Organic biomass – 2.57
Photo fermentation Solar Organic biomass – 2.83
Solar photovoltaic electrolysis Solar Water 12.0 to 54.5 5.78 to 23.27
Solar thermal electrolysis Solar Water 22.1 to 421.0 5.10 to 10.49
Wind electrolysis Wind Water 499.6 to 504.8 5.89 to 6.03
Nuclear electrolysis Nuclear Water – 4.15 to 7.00
Nuclear thermolysis Nuclear Water 39.6 to 2107.6 2.17 to 2.63
Solar thermolysis Solar Water 5.7 to 16.0 7.89 to 8.40
Photoelectrolysis Solar Water – 10.36

9
B. Zhang et al. Journal of Electronic Science and Technology 19 (2021) 100080

Table 2
Evaluation indexes of several hydrogen-production technologies.
No. Hydrogen-production Equivalent greenhouse gas emission [(kg, CO2)/(kg, Energy consumption [MJ/(kg, Life cycle cost [CNY/(kg,
technology H2)] H2)] H2)]

1 Biomass gasification 400 to 5600 4 to 20 9.7 to 22.2


2 Photovoltaic technology 2400 to 6800 30 to 80 36.6 to 61.3
3 Wind power 600 to 970 5 to 12 22.3 to 59.8
4 Nuclear energy/ 300 to 860 360 to 410 12.8 to 36.9
thermochemistry
5 Natural gas 3900 to 12900 165 to 360 10.4 to 27.6
6 Coal gasification 5000 to 11300 190 to 325 8.3 to 19.5

4.2. Special requirements

Although more than 90% of the global hydrogen is manufactured from well-developed, low-cost methane and other gaseous fossil
hydrocarbon fuels, inevitably, the subsequent compression, storage, transportation, and distribution of the technology will cause a
steady increase in the selling price to consumers. Moreover, many basic supporting facilities are inadequate, and the technologies cannot
meet the requirements for small-scale distribution of hydrogen sources in power supply and other fields, including transportation,
military submarines, and individual combat at the present stage [105,106].
These fields often require onsite hydrogen production or conversion for convenience and safety. Fig. 8 lists the existing production
methods from the point of onsite and offsite. Among them, the conversion of methanol to hydrogen has attracted significant attention
because of its characteristics of liquid carrier, high energy density, safety and reliability, low storage and transportation costs, relatively
mild conversion conditions for hydrogen production, no sulfur, low toxicity, and relatively easy realization of the technical process
[107]. Conversely, methanol can be derived from either fossils or renewable sources. Presently, the major research direction of this
technology is toward the replacement for precious metal catalysts, which are expensive and susceptible to poisoning. One of the
developed candidates is Cu-based catalysts, which are inexpensive and exhibit outstanding activity at low temperatures. The research
mostly addresses the structural composition, spatial distribution, coating form, etc., in the determination of the methanol conversion,
reaction rate, and hydrogen yields.

4.3. Process coupling/enhancement

Since most of the present hydrogen-production technologies have the problems of high energy consumption, severe pollution, poor
efficiency, and low purity of hydrogen products, the chemical process enhancement is frequently proposed for various renewable
sources or sustainable technologies. The former case has been reviewed in the previous section of this paper, whereas the latter case has
been exploited by coupling with sonochemistry and sonoelectrochemistry [108], geothermal energy [109], microwave heating [110],
carbon-assisted water electrolysis [111], etc.
Another direction is to integrate multiple processes, e.g., reaction and separation. This could overcome most of the existing problems
while reducing the investment and preparation costs and simplifying the technical processes. Such processes include photovoltaic-
thermoelectric systems [112], sorption-enhanced carbonate looping processes [113], membrane coupling with thermochemical pro-
cesses [114], and membrane-assisted gas switching reforming [115]. For example, previously, our team developed a kind of catalytic
carbon membrane that exhibits outstanding properties for improving the methanol conversion and hydrogen yields during the steam
reforming of methanol by integrating the reaction with a membrane separation process [116].
In summary, all the abovementioned protocols have been verified to be effective for improving the efficiency of the current
hydrogen-production technologies.

5. Outlook and remarks

Hydrogen has been recognized as the cleanest and most promising energy source in the 21st century. Additionally, it is an important
chemical raw material widely used in industries. Presently, a large variety of feasible techniques have been developed for hydrogen
production, including reforming/oxidation/gasification of fossil fuels, decomposition of hydrogen-containing resources, water elec-
trolysis, microbial fermentation, and PC [117]. However, most of them have several drawbacks, including significant energy con-
sumption, serious pollution, inferior conversion efficiency, and high content of impurities in products. Therefore, the following aspects
should be considered in future research to conquer the confronted issues.

1) Theoretical research. To achieve substantial progress in theory, the interface engineering technology can be applied to investigate
the steam reforming of hydrocarbons from the microscopic perspective of atomic engineering methods [118]. Conversely, computer
modeling and calculations are also helpful for predicting, evaluating, and optimizing different types of hydrogen-production
methods [119]. This aspect can guide the development of experimental routes and the optimization of process conditions,
thereby improving the efficiency and reducing the hydrogen-production costs.
2) Development of raw materials. The production cost would be greatly reduced by fully exploiting renewable resources, biomass, and
even waste as raw materials. This will certainly result in significant economic, social, and environmental benefits. In fact, several

10
B. Zhang et al. Journal of Electronic Science and Technology 19 (2021) 100080

Fig. 8. Hydrogen-production methods with respect to onsite and offsite.

attempts have been made to utilize such raw materials to produce hydrogen, including biomass and residual waste [120], plastic
waste [121], and seawater [122]. These studies inspired our new ideas and directions.
3) Process coupling and sustainable energy use. Besides the numerous raw materials, current hydrogen-production technologies are
associated with complicated processes and high energy consumption. Therefore, the process coupling strategy is recommended for
combining some essential processes to enhance the efficiency of hydrogen-production technologies, e.g., PC-electrolysis hybrid
systems [123]. Moreover, the use of sustainable energy is recommended, including solar energy, geothermal energy, nuclear power,
and wind power [124]. These measures will significantly improve the efficiency of hydrogen-production technologies and opera-
tional flexibility and convenience, as well as the future market competition and application prospects.

Funds

This work was supported by the National Natural Science Foundation of China under Grant No. 20906063; the Liaoning BaiQianWan
Talents Program under Grant No. 2018921046; the Scientific Research Project of Liaoning Provincial Department of Education under
Grant No. LJGD2020002; the Shenyang Youth Science and Technology Project under Grant No. RC200325.

Declaration of competing interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.

References

[1] S.H. Mohr, J. Wang, G. Ellem, J. Ward, D. Giurco, Projection of world fossil fuels by country, Fuel 141 (Feb. 2015) 120–135.
[2] X.-L. Ouyang, B.-Q. Lin, Impacts of increasing renewable energy subsidies and phasing out fossil fuel subsidies in China, Renew. Sustain. Energy Rev. 37 (Sept.
2014) 933–942.
[3] V.S. Arutyunov, G.V. Lisichkin, Energy resources of the 21st century: problems and forecasts. Can renewable energy sources replace fossil fuels? Russ. Chem.
Rev. 86 (8) (Aug. 2017) 777–804.
[4] J.M. Sanchez, M.M. Barreiro, M. Maro~ no, Bench-scale study of separation of hydrogen from gasification gases using a palladium-based membrane reactor, Fuel
116 (Jan. 2014) 894–903.
[5] C.A. Pereira, P.M. Coelho, J.F. Fernandes, M.H. Gomes, “Study of an energy mix for the production of hydrogen,” Intl, Journal of Hydrogen Energy 42 (2) (Jan.
2017) 1375–1382.
[6] T. da Silva Veras, T.S. Mozer, D. da Costa Rubim Messeder dos Santos, A. da Silva Cesar, Hydrogen: trends, production and characterization of the main process
worldwide, Intl. Journal of Hydrogen Energy 42 (4) (Jan. 2017) 2018–2033.
[7] S. Suman, Hybrid nuclear-renewable energy systems: a review, J. Clean. Prod. 181 (Apr. 2018) 166–177.
[8] B. Parkinson, P. Balcombe, J.F. Speirs, A.D. Hawkes, K. Hellgardt, Levelized cost of CO2 mitigation from hydrogen production routes, Energy Environ. Sci. 12 (1)
(2019) 19–40.
[9] W.-H. Chen, C.-Y. Chen, Water gas shift reaction for hydrogen production and carbon dioxide capture: a review, Appl. Energy 258 (Jan. 2020) 1–25, 114078.
[10] D. Damma, P.G. Smirniotis, Recent advances in iron-based high-temperature water-gas shift catalysis for hydrogen production, Current Opinion in Chemical
Engineering 21 (Sept. 2018) 103–110.
[11] A.I. Osman, Catalytic hydrogen production from methane partial oxidation: mechanism and kinetic study, Chem. Eng. Technol. 43 (4) (Apr. 2020) 641–648.
[12] Y.-C. Gao, J.-G. Jiang, Y. Meng, F. Yan, A. Aihemaiti, A review of recent developments in hydrogen production via biogas dry reforming, Energy Convers.
Manag. 171 (Sept. 2018) 133–155.
[13] A.F.S. Muhammad, A. Awad, R. Saidur, N. Masiran, A. Salam, B. Abdullah, Recent advances in cleaner hydrogen productions via thermo-catalytic decomposition
of methane: admixture with hydrocarbon, Intl. Journal of Hydrogen Energy 43 (41) (Oct. 2018) 18713–18734.
[14] M. Taghizadeh, F. Aghili, Recent advances in membrane reactors for hydrogen production by steam reforming of ethanol as a renewable resource, Rev. Chem.
Eng. 35 (3) (Mar. 2019) 377–392.
[15] S. Ogo, Y. Sekine, Low temperature hydrogen production by ethanol steam reforming over supported metal catalysts in an electric field, J. Jpn. Petrol. Inst. 62
(6) (2019) 264–271.
[16] M.A. Salam, K. Ahmed, N. Akter, T. Hossain, B. Abdullah, A review of hydrogen production via biomass gasification and its prospect in Bangladesh, Intl. Journal
of Hydrogen Energy 43 (32) (Aug. 2018) 14944–14973.
[17] M. Yasuda, T. Matsumoto, T. Yamashita, Sacrificial hydrogen production over TiO2-based photocatalysts: polyols, carboxylic acids, and saccharides, Renew.
Sustain. Energy Rev. 81 (Jan. 2018) 1627–1635.

11
B. Zhang et al. Journal of Electronic Science and Technology 19 (2021) 100080

[18] B.-L. Dou, H. Zhang, Y.-C. Song, et al., Hydrogen production from the thermochemical conversion of biomass: issues and challenges, Sustainable Energy & Fuels
3 (2) (2019) 314–342.
[19] M. Anwar, S.-L. Lou, L. Chen, H. Li, Z.-L. Hu, Recent advancement and strategy on bio-hydrogen production from photosynthetic microalgae, Bioresour.
Technol. 292 (Nov. 2019) 1–13, 121972.
[20] S.-J. Wang, T. Zhang, M.-D. Bao, H.-J. Su, P. Xu, Microbial production of hydrogen by mixed culture technologies: a review, Biotechnol. J. 15 (1) (Jan. 2020),
1900297:1-8.
[21] R.S. Tan, T.A.T. Abdullah, A. Johari, K.M. Isa, Catalytic steam reforming of tar for enhancing hydrogen production from biomass gasification: a review, Front.
Energy 14 (3) (Mar. 2020) 545–569.
[22] H. Song, H.S. Jung, S. Uhm, Recent progress for hydrogen production from biogas and its effective applications, Applied Chemistry for Engineering 31 (1)
(2020) 1–6.
[23] N. Sanchez, R. Ruiz, V. Hacker, M. Cobo, “Impact of bioethanol impurities on steam reforming for hydrogen production: a review,” Intl, Journal of Hydrogen
Energy 45 (21) (Apr. 2020) 11923–11942.
[24] A.B.A. Ibrahim, H. Akilli, Supercritical water gasification of wastewater sludge for hydrogen production, Intl. Journal of Hydrogen Energy 44 (21) (Apr. 2019)
10328–10349.
[25] A. Fasolini, R. Cucciniello, E. Paone, F. Mauriello, T. Tabanelli, A short overview on the hydrogen production via aqueous phase reforming (APR) of cellulose,
C6-C5 sugars and polyols, Catalysts 9 (11) (Nov. 2019) 1–16, 917.
[26] H.J. Park, S.Y. Lee, T.K. Lee, H.J. Kim, Y.M. Lee, N3-butyl imidazolium-based anion exchange membranes blended with poly(vinyl alcohol) for alkaline water
electrolysis, J. Membr. Sci. 611 (Oct. 2020) 1–11, 118355.
[27] J. Chi, H.-M. Yu, Water electrolysis based on renewable energy for hydrogen production, Chin. J. Catal. 39 (3) (Mar. 2018) 390–394.
[28] Y.-C. Huang, L. Hu, R. Liu, et al., Nitrogen treatment generates tunable nanohybridization of Ni5P4 nanosheets with nickel hydr(oxy)oxides for efficient
hydrogen production in alkaline, seawater and acidic media, Appl. Catal. B Environ. 251 (Aug. 2019) 181–194.
[29] Y.-X. Gao, T.-Z. Xiong, Y. Li, Y.-C. Huang, Y.-P. Li, M.S.J.T. Balogun, A simple and scalable approach to remarkably boost the overall water splitting activity of
stainless steel electrocatalysts, ACS Omega 4 (14) (Sept. 2019) 16130–16138.
[30] I. Vincent, D. Bessarabov, Low cost hydrogen production by anion exchange membrane electrolysis: a review, Renew. Sustain. Energy Rev. 81 (Jan. 2018)
1690–1704.
[31] Y.-W. Hu, D. Huang, J.-N. Zhang, Y.-C. Huang, M.S.J.T. Balogun, Y.-X. Tong, Dual doping induced interfacial engineering of Fe2N/Fe3N hybrids with favorable
d-band towards efficient overall water splitting, ChemCatChem 11 (24) (Dec. 2019) 6051–6060.
[32] P. Mishra, S. Krishnan, S. Rana, L. Singh, M. Sakinah, Z.A. Wahid, Outlook of fermentative hydrogen production techniques: an overview of dark, photo and
integrated dark-photo fermentative approach to biomass, Energy Strategy Reviews 24 (Apr. 2019) 27–37.
[33] A.H.S. Hassan, T. Mietzel, R. Brunstermann, et al., Fermentative hydrogen production from low-value substrates, World J. Microbiol. Biotechnol. 34 (Nov. 2018)
1–11, 176.
[34] K. Chandrasekhar, Y.J. Lee, D.W. Lee, “Biohydrogen production: strategies to improve process efficiency through microbial routes,” Intl, J. Mol. Sci. 16 (4) (Apr.
2015) 8266–8293.
[35] Y. Wang, H.-L. Yang, X.-N. Zhang, F. Han, W.-F. Tu, W.-Q. Yang, Microalgal hydrogen production, Small Methods 4 (3) (Mar. 2020) 1900514, https://doi.org/
10.1002/smtd.201900514.
[36] S. Gao, W.-H. Fan, Y. Liu, D.-Y. Jiang, Q. Duan, “Artificial water-soluble systems inspired by[FeFe]-hydrogenases for electro- and photocatalytic hydrogen
production,” Intl, Journal of Hydrogen Energy 45 (7) (Feb. 2020) 4305–4327.
[37] B.M. Barney, Aerobic nitrogen-fixing bacteria for hydrogen and ammonium production: current state and perspectives, Appl. Microbiol. Biotechnol. 104 (4)
(Feb. 2020) 1383–1399.
[38] A. Valle, D. Cantero, J. Bolívar, Metabolic engineering for the optimization of hydrogen production in Escherichia coli: a review, Biotechnol. Adv. 37 (5) (Sept.-
Oct. 2019) 616–633.
[39] A. Krishnan, X. Qian, G. Ananyev, D.-S. Lun, G.C. Dismukes, Rewiring of cyanobacterial metabolism for hydrogen production: synthetic biology approaches and
challenges, in: W.-W. Zhang, X.-Y. Song (Eds.), In Synthetic Biology Of Cyanobacteria, Springer, Singapore, 2018, pp. 171–213.
[40] A. Latifi, L. Avilan, M. Brugna, Clostridial whole cell and enzyme systems for hydrogen production: current state and perspectives, Appl. Microbiol. Biotechnol.
103 (2) (Jan. 2019) 567–575.
[41] K. Chandrasekhar, S. Kumar, B.D. Lee, S.H. Kim, Waste based hydrogen production for circular bioeconomy: current status and future directions, Bioresour.
Technol. 302 (Apr. 2020) 1–10, 122920.
[42] T. Jarunglumlert, C. Prommuak, N. Putmai, P. Pavasant, “Scaling-up bio-hydrogen production from food waste: feasibilities and challenges,” Intl, Journal of
Hydrogen Energy 43 (2) (Jan. 2018) 634–648.
[43] G. Sołowski, M.S. Shalaby, H. Abdallah, A.M. Shaban, A. Cenian, Production of hydrogen from biomass and its separation using membrane technology, Renew.
Sustain. Energy Rev. 82 (Feb. 2018) 3152–3167.
[44] S.K.S. Patel, J.K. Lee, V.C. Kalia, Nanoparticles in biological hydrogen production: an overview, Indian J. Microbiol. 58 (1) (Mar. 2018) 8–18.
[45] R. Łukajtis, I. Hołowacz, K. Kucharska, et al., Hydrogen production from biomass using dark fermentation, Renew. Sustain. Energy Rev. 91 (Aug. 2018)
665–694.
[46] Y. Liu, R.-J. Lin, Y. Man, J.-Z. Ren, Recent developments of hydrogen production from sewage sludge by biological and thermochemical process, Intl. Journal of
Hydrogen Energy 44 (36) (Jul. 2019) 19676–19697.
[47] S.H. Kim, A. Mudhoo, A. Pugazhendhi, et al., A perspective on galactose-based fermentative hydrogen production from macroalgal biomass: trends and
opportunities, Bioresour. Technol. 280 (May 2019) 447–458.
[48] Z.M.A. Bundhoo, “Potential of bio-hydrogen production from dark fermentation of crop residues: a review,” Intl, Journal of Hydrogen Energy 44 (32) (Jun.
2019) 17346–17362.
[49] V.L. Pachapur, P. Kutty, P. Pachapur, et al., Seed pretreatment for increased hydrogen production using mixed-culture systems with advantages over pure-
culture systems, Energies 12 (3) (Feb. 2019) 1–26, 530.
[50] G. Yang, J.-L. Wang, Various additives for improving dark fermentative hydrogen production: a review, Renew. Sustain. Energy Rev. 95 (Nov. 2018) 130–146.
[51] H. Wu, H.-L. Tan, C.Y. Toe, et al., Photocatalytic and photoelectrochemical systems: similarities and differences, Adv. Mater. 32 (18) (May 2020) 1–21,
1904717.
[52] J.H. Kim, D. Hansora, P. Sharma, J.-W. Jang, J.S. Lee, “Toward practical solar hydrogen production—an artificial photosynthetic leaf-to-farm challenge, Chem.
Soc. Rev. 48 (7) (Mar. 2019) 1908–1971.
[53] X.-L. Huang, X. Tong, Z.-M. Wang, Rational design of colloidal core/shell quantum dots for optoelectronic applications, Journal of Electronic Science and
Technology 18 (2) (Jun. 2020) 105–118.
[54] R.-G. Li, Latest progress in hydrogen production from solar water splitting via photocatalysis, photoelectrochemical, and photovoltaic-photoelectrochemical
solutions, Chin. J. Catal. 38 (1) (Jan. 2017) 5–12.
[55] S. Bellani, M.R. Antognazza, F. Bonaccorso, Carbon-based photocathode materials for solar hydrogen production, Adv. Mater. 31 (9) (Mar. 2019), 1801446:1-
33.
[56] M. Fereidooni, A. Mostafaeipour, V. Kalantar, H. Goudarzi, A comprehensive evaluation of hydrogen production from photovoltaic power station, Renew.
Sustain. Energy Rev. 82 (Feb. 2018) 415–423.
[57] M. Ahmed, I. Dincer, A review on photoelectrochemical hydrogen production systems: challenges and future directions, Intl. Journal of Hydrogen Energy 44 (5)
(Jan. 2019) 2474–2507.
[58] F. Varas-Concha, D. Guzman, M. Isaacs, C. Saez-Navarrete, Operational conditions affecting hydrogen production by the photoreforming of organic compounds
using titania nanoparticles with gold, Energy Technol. 6 (2) (Feb. 2018) 416–431.

12
B. Zhang et al. Journal of Electronic Science and Technology 19 (2021) 100080

[59] L.-L. Ji, C.-C. Lyu, Z.-F. Chen, Z.-P. Huang, C. Zhang, Nickel-based (photo)electrocatalysts for hydrogen production, Adv. Mater. 30 (7) (Apr. 2018) 1–7,
1705653.
[60] Y. Liu, D.-L. Huang, M. Cheng, et al., Metal sulfide/MOF-based composites as visible-light-driven photocatalysts for enhanced hydrogen production from water
splitting, Coord. Chem. Rev. 409 (May 2020) 213220, 1-19.
[61] C. Wu, J. Zhang, X. Tong, et al., A critical review on enhancement of photocatalytic hydrogen production by molybdenum disulfide: from growth to interfacial
activities, Small 15 (35) (Aug. 2019) 1–25, 1900578.
[62] A. Savateev, M. Antonietti, Ionic carbon nitrides in solar hydrogen production and organic synthesis: exciting chemistry and economic advantages,
ChemCatChem 11 (24) (Dec. 2019) 6166–6176.
[63] J. Kampmann, S. Betzler, H. Hajiyani, et al., How photocorrosion can trick you: a detailed study on low-bandgap Li doped CuO photocathodes for solar
hydrogen production, Nanoscale 12 (14) (Mar. 2020) 7766–7775.
[64] W.-T. Qiu, Y.-C. Huang, S.-T. Tang, H.-B. Ji, Y.-X. Tong, Thin-layer indium oxide and cobalt oxyhydroxide cobalt-modified BiVO4 photoanode for solar-assisted
water electrolysis, J. Phys. Chem. C 121 (32) (Jul. 2017) 17150–17159.
[65] H. Yi, D.-L. Huang, L. Qin, et al., Selective prepared carbon nanomaterials for advanced photocatalytic application in environmental pollutant treatment and
hydrogen production, Appl. Catal. B Environ. 239 (Dec. 2018) 408–424.
[66] B.-J. Ma, D.-K. Li, X.-Y. Wang, K.-Y. Lin, Molybdenum-based co-catalysts in photocatalytic hydrogen production: categories, structures, and roles, ChemSusChem
11 (22) (Nov. 2018) 3871–3881.
[67] R.S. Sprick, J.-X. Jiang, B. Bonillo, et al., Tunable organic photocatalysts for visible-light-driven hydrogen evolution, J. Am. Chem. Soc. 137 (9) (Feb. 2015)
3265–3270.
[68] J. Kosco, M. Bidwell, H. Cha, et al., Enhanced photocatalytic hydrogen evolution from organic semiconductor heterojunction nanoparticles, Nat. Mater. 19 (5)
(Feb. 2020) 559–565.
[69] J. Chen, X.-P. Tao, L. Tao, et al., Novel conjugated organic polymers as candidates for visible-light-driven photocatalytic hydrogen production, Appl. Catal. B
Environ. 241 (Feb. 2019) 461–470.
[70] J. Corredor, M.J. Rivero, C.M. Rangel, F. Gloaguen, I. Ortiz, Comprehensive review and future perspectives on the photocatalytic hydrogen production, J. Chem.
Technol. Biotechnol. 94 (10) (Oct. 2019) 3049–3063.
[71] T. Seadira, G. Sadanandam, T.A. Ntho, X.-J. Lu, C.M. Masuku, M. Scurrell, Hydrogen production from glycerol reforming: conventional and green production,
Rev. Chem. Eng. 34 (5) (Aug. 2018) 695–726.
[72] A. Kumar, R. Singh, A.S.K. Sinha, Catalyst modification strategies to enhance the catalyst activity and stability during steam reforming of acetic acid for
hydrogen production, Intl. Journal of Hydrogen Energy 44 (26) (May 2019) 12983–13010.
[73] M. Navlani-Garcia, D. Salinas-Torres, D. Cazorla-Amoros, Hydrogen production from formic acid attained by bimetallic heterogeneous PdAg catalytic systems,
Energies 12 (21) (Oct. 2019) 4027, 1-27.
[74] M. Navlani-García, K. Mori, D. Salinas-Torres, Y. Kuwahara, H. Yamashita, New approaches toward the hydrogen production from formic acid dehydrogenation
over Pd-based heterogeneous catalysts, Frontiers in Materials 6 (Mar. 2019) 1–18, 44.
[75] X. Wang, Q.-L. Meng, L.-Q. Gao, et al., Recent progress in hydrogen production from formic acid decomposition, Intl. Journal of Hydrogen Energy 43 (14) (Apr.
2018) 7055–7071.
[76] S. Díaz-Abad, M. Mill an, M.A. Rodrigo, J. Lobato, “Review of anodic catalysts for SO2 depolarized electrolysis for ‘green hydrogen’ production, Catalysts 9 (1)
(Jan. 2019) 63, 1-16.
[77] Y.-X. Cheng, X. Wu, H.-L. Xu, Catalytic decomposition of hydrous hydrazine for hydrogen production, Sustainable Energy & Fuels 3 (2) (Nov. 2019) 343–365.
[78] M. Navlani-García, K. Mori, Y. Kuwahara, H. Yamashita, Recent strategies targeting efficient hydrogen production from chemical hydrogen storage materials
over carbon-supported catalysts, NPG Asia Mater. 10 (4) (Apr. 2018) 277–292.
[79] A. Srifa, K. Okura, T. Okanishi, H. Muroyama, T. Matsui, K. Eguchi, Hydrogen production by ammonia decomposition over Cs-modified Co3Mo3N catalysts,
Appl. Catal. B Environ. 218 (Dec. 2017) 1–8.
[80] I. Dincer, C. Acar, Review and evaluation of hydrogen production methods for better sustainability, Intl. Journal of Hydrogen Energy 40 (34) (Sept. 2015)
11094–11111.
[81] S.T. Revankar, Nuclear hydrogen production, in: H. Bindra, S. Revankar (Eds.), In Storage And Hybridization Of Nuclear Energy, Academic Press, Amsterdam,
2019, pp. 49–117, ch. 4.
[82] M. Kayfeci, A. Keçebaş, M. Bayat, in: F. Calise, M.D. D’Accadia, M. Santarelli, A. Lanzini, D. Ferrero (Eds.), “Hydrogen Production, ” in Solar Hydrogen Production,
Academic Press, Amsterdam, 2019, pp. 45–83, ch. 3.
[83] P. Zhang, L.-J. Wang, S.-Z. Chen, J.-M. Xu, Progress of nuclear hydrogen production through the iodine-sulfur process in China, Renew. Sustain. Energy Rev. 81
(Jan. 2018) 1802–1812.
[84] Q. Sun, Q.-X. Gao, P. Zhang, W. Peng, S.-Z. Chen, Modeling sulfuric acid decomposition in a bayonet heat exchanger in the iodine-sulfur cycle for hydrogen
production, Appl. Energy 277 (Nov. 2020) 1–12, 115611.
[85] R.D. Solunke, G. Veser, Hydrogen production via chemical looping steam reforming in a periodically operated fixed-bed reactor, Ind. Eng. Chem. Res. 49 (21)
(Jun. 2010) 11037–11044.
[86] X.-H. Wang, X.-C. Du, W.-B. Yu, J.-S. Zhang, J.-J. Wei, Coproduction of hydrogen and methanol from methane by chemical looping reforming, Ind. Eng. Chem.
Res. 58 (24) (May 2019) 10296–10306.
[87] M.-J. Wang, G.-Z. Wang, Z.-X. Sun, Y.-K. Zhang, D. Xu, Review of renewable energy-based hydrogen production processes for sustainable energy innovation,
Global Energy Interconnection 2 (5) (Oct. 2019) 436–443.
[88] C.W. Hsu, C.M. Tung, C.-Y. Lin, Industrialization roadmap model for fermentative hydrogen production from biomass in Taiwan, Intl. Journal of Hydrogen
Energy 42 (45) (Nov. 2017) 27460–27470.
[89] B. Zohuri, Large-scale hydrogen production, in: B. Zohuri (Ed.), In Hydrogen Energy: Challenges And Solutions For a Cleaner Future, Springer International
Publishing, Cham, 2019, pp. 229–255.
[90] M.C.J. Bradford, M.V. Konduru, D.X. Fuentes, Preparation, characterization and application of Cr2O3/ZnO catalysts for methanol synthesis, Fuel Process.
Technol. 83 (1–3) (Sept. 2003) 11–25.
[91] F. Dalena, A. Senatore, A. Marino, A. Gordano, M. Basile, A. Basile, in: A. Basile, F. Dalena (Eds.), “Methanol Production and Applications: an Overview,” in
Methanol, Elsevier, Amsterdam, 2018, pp. 3–28, ch. 1.
[92] D. Sheldon, “Methanol production—a technical history, Johnson Matthey Technology Review 61 (3) (Jul. 2017) 172–182.
[93] W. Su, P. Liu, C.-Q. Cai, et al., Hydrogen production and heavy metal immobilization using hyperaccumulators in supercritical water gasification, J. Hazard
Mater. 402 (Jan. 2021) 1–9, 123541.
[94] N.J. Wagner, M. Coertzen, R.H. Matjie, J.C. van Dyk, Coal gasification, in: I. Suarez-Ruiz, J.C. Crelling (Eds.), In Applied Coal Petrology, Elsevier, Amsterdam,
2008, pp. 119–144, ch. 5.
[95] J. Hn at, M. Paidar, K. Bouzek, Hydrogen production by electrolysis, in: A. Iulianelli, A. Basile (Eds.), In Current Trends And Future Developments On (Bio-)
Membranes, Elsevier, Amsterdam, 2020, pp. 91–117.
[96] F.-F. Cai, J.J. Ibrahim, Y. Fu, W. Kong, J. Zhang, Y.-H. Sun, Low-temperature hydrogen production from methanol steam reforming on Zn-modified Pt/MoC
catalysts, Appl. Catal. B Environ. 264 (May 2020) 1–13, 118500.
[97] D.R. Palo, R.A. Dagle, J.D. Holladay, Methanol steam reforming for hydrogen production, Chem. Rev. 107 (10) (Sept. 2007) 3992–4021.
[98] M.A. Khan, I. Al-Shankiti, A. Ziani, N. Wehbe, H. Idriss, A stable integrated photoelectrochemical reactor for H2 production from water attains a solar-to-
hydrogen efficiency of 18% at 15 suns and 13% at 207 suns, Angew. Chem. 132 (35) (Aug. 2020) 14912–14918.
[99] Office of Energy Efficiency & Renewable Energy, Hydrogen production. [Online]. Available:. https://www.energy.gov/eere/fuelcells/hydrogen-production.

13
B. Zhang et al. Journal of Electronic Science and Technology 19 (2021) 100080

[100] S.E. Hosseini, M.A. Wahid, Hydrogen production from renewable and sustainable energy resources: promising green energy carrier for clean development,
Renew. Sustain. Energy Rev. 57 (May 2016) 850–866.
[101] H.-B. Gao, W.-L. Zhen, J.-T. Ma, G.-X. Lu, High efficient solar hydrogen generation by modulation of Co-Ni sulfide (220) surface structure and adjusting
adsorption hydrogen energy, Appl. Catal. B Environ. 206 (Jun. 2017) 353–363.
[102] F. Dawood, M. Anda, G.M. Shafiullah, “Hydrogen production for energy: an overview,” Intl, Journal of Hydrogen Energy 45 (7) (Feb. 2020) 3847–3869.

[103] R.S. El-Emam, H. Ozcan, Comprehensive review on the techno-economics of sustainable large-scale clean hydrogen production, J. Clean. Prod. 220 (May 2019)
593–609.
[104] X.-S. Xie, W.-J. Yang, W. Shi, S.-S. Zhang, Z.-H. Wang, J.-H. Zhou, “Life cycle assessment of technologies for hydrogen production—a review, Chem. Ind. Eng.
Prog. 37 (6) (Jun. 2018) 2147–2158.
[105] B.D. Solomon, A. Banerjee, A global survey of hydrogen energy research, development and policy, Energy Pol. 34 (7) (May 2006) 781–792.
[106] F. Zhang, P.-C. Zhao, M. Niu, J. Maddy, “The survey of key technologies in hydrogen energy storage,” Intl, Journal of Hydrogen Energy 41 (33) (Sept. 2016)
14535–14552.
[107] B. Zhang, D.-D. Zhao, Y.-H. Wu, H.-J. Liu, T.-H. Wang, J.-S. Qiu, Fabrication and application of catalytic carbon membranes for hydrogen production from
methanol steam reforming, Ind. Eng. Chem. Res. 54 (2) (Dec. 2015) 623–632.
[108] M.H. Islam, O.S. Burheim, B.G. Pollet, Sonochemical and sonoelectrochemical production of hydrogen, Ultrason. Sonochem. 51 (Mar. 2019) 533–555.
[109] M. Ghazvini, M. Sadeghzadeh, M.H. Ahmadi, S. Moosavi, F. Pourfayaz, “Geothermal energy use in hydrogen production: a review,” Intl, Journal of Energy
Research 43 (14) (Nov. 2019) 7823–7851.
[110] H. Fukushima, Hydrogen production by microwave steam reforming, J. Jpn. Petrol. Inst. 61 (2) (2018) 106–112.
[111] H. Ju, S. Badwal, S. Giddey, A comprehensive review of carbon and hydrocarbon assisted water electrolysis for hydrogen production, Appl. Energy 231 (Dec.
2018) 502–533.
[112] J.-P. Xu, Q.-L. Li, H.-P. Xie, T. Ni, C. Ouyang, Tech-integrated paradigm based approaches towards carbon-free hydrogen production, Renew. Sustain. Energy
Rev. 82 (Feb. 2018) 4279–4295.
[113] D.P. Hanak, S. Michalski, V. Manovic, From post-combustion carbon capture to sorption-enhanced hydrogen production: a state-of-the-art review of carbonate
looping process feasibility, Energy Convers. Manag. 177 (Dec. 2018) 428–452.
[114] G.-Z. Ji, J.-G. Yao, P.T. Clough, et al., Enhanced hydrogen production from thermochemical processes, Energy Environ. Sci. 11 (10) (2018) 2647–2672.
[115] S.A. Wassie, J.A. Medrano, A. Zaabout, et al., Hydrogen production with integrated CO2 capture in a membrane assisted gas switching reforming reactor: proof-
of-concept, Intl. Journal of Hydrogen Energy 43 (12) (Mar. 2018) 6177–6190.
[116] B. Zhang, D. Wang, J.-L. Zhou, et al., Process intensification of the hydrogen production reaction using a carbon membrane reactor: kinetics analysis, Energy
Technol. 5 (11) (Nov. 2017) 1990–1997.
[117] K.I. Fujita, Development of efficient methods for organic synthesis, hydrogen storage, and hydrogen production based on catalytic dehydrogenation of organic
molecules, Journal of Synthetic Organic Chemistry Japan 77 (2) (2019) 112–119.
[118] S. Chen, C.-L. Pei, J.-L. Gong, Insights into interface engineering in steam reforming reactions for hydrogen production, Energy Environ. Sci. 12 (12) (Oct. 2019)
3473–3495.
[119] S.F. Ardabili, B. Najafi, S. Shamshirband, B.M. Bidgoli, R.C. Deo, K.W. Chau, Computational intelligence approach for modeling hydrogen production: a review,
Engineering Applications of Computational Fluid Mechanics 12 (1) (Mar. 2018) 438–458.
[120] M. Shahabuddin, B.B. Krishna, T. Bhaskar, G. Perkins, Advances in the thermo-chemical production of hydrogen from biomass and residual wastes: summary of
recent techno-economic analyses, Bioresour. Technol. 299 (Mar. 2020) 1–12, 122557.
[121] S.S. Sharma, V.S. Batra, Production of hydrogen and carbon nanotubes via catalytic thermo-chemical conversion of plastic waste: Review, J. Chem. Technol.
Biotechnol. 95 (1) (Jan. 2020) 11–19.
[122] V. Kumaravel, A. Abdel-Wahab, A short review on hydrogen, biofuel, and electricity production using seawater as a medium, Energy & Fuels 32 (6) (May 2018)
6423–6437.
[123] Y. Miseki, K. Sayama, Photocatalytic water splitting for solar hydrogen production using the carbonate effect and the Z-scheme reaction, Advanced Energy
Materials 9 (23) (Jun. 2019) 1–15, 1801294.
[124] Z. Li, P. Guo, R.-H. Han, H.-X. Sun, Current status and development trend of wind power generation-based hydrogen production technology, Energy Explor.
Exploit. 37 (1) (2019) 5–25.

Bing Zhang received his Ph.D. degree in chemical technology from Dalian University of Technology, Dalian in 2007. Now he is a full professor
with the School of Petrochemical Engineering, Shenyang University of Technology, Liaoyang. He has published more than 80 journal papers
and 30 conference papers. His main research interests include the porous materials fabrication, membrane filtration, gas separation,
wastewater treatment, hydrogen production, and membrane catalysis.

Sui-Xin Zhang received his B.S. degree from Hefei University of Technology, Hefei in 2019. He is currently pursuing his M.S. degree in
chemical engineering with Shenyang University of Technology. His research interest is mainly focused on the preparation of catalytic
membranes for enhanced hydrogen production from steam reforming of methanol.

14
B. Zhang et al. Journal of Electronic Science and Technology 19 (2021) 100080

Rui Yao received his B.S. degree from Shenyang University of Technology in 2018. Now he is pursuing his M.S. degree with the School of
Petrochemical Engineering, Shenyang University of Technology. He is working on the fabrication and applications of membrane materials for
environmental protection.

Yong-Hong Wu obtained her B.S. degree from Shenyang University of Chemical Technology, Shenyang in 2002. Then, she worked as a
research engineer in an adhesive company for new product development. In 2007, she joined Prof. Bing Zhang’s team as a researcher with
Shenyang University of Technology. Now she is a lecture with the School of Petrochemical Engineering, Shenyang University of Technology.
Her research interests include membrane production, gas purification, oily wastewater treatment, and hydrogen production.

Jie-Shan Qiu received his Ph.D. degree in organic chemical engineering from Dalian University of Technology in 1990. Now he is the
Changjiang Distinguished Professor in carbon science and chemical engineering, and the Dean of the College of Chemical Engineering, Beijing
University of Chemical Technology, Beijing. His research encompasses both fundamental and applied aspects of carbon materials and science,
focusing on the methodologies for producing carbon materials from coal and coal-derived byproducts, such as coal tar pitch and their ap-
plications in energy storage and conversion (solar cells, supercapacitors, batteries, etc.), catalysis, and environment protection.

15

View publication stats

You might also like