You are on page 1of 16

Biomass and Bioenergy 144 (2021) 105920

Contents lists available at ScienceDirect

Biomass and Bioenergy


journal homepage: http://www.elsevier.com/locate/biombioe

Biomass-to-hydrogen: A review of main routes production, processes


evaluation and techno-economical assessment
Thibaut Lepage, Maroua Kammoun, Quentin Schmetz, Aurore Richel *
University of Liège – Gembloux Agro-Bio Tech, Biomass and Green Technologies, Passage des Déportés, 2, B-5030, Gembloux, Belgium

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

Keywords: Hydrogen is viewed as a sustainable strategic alternative to fossil fuels, especially in the field of road and air
Bioenergy transport. Currently, hydrogen production is derived from fossil fuels or is manufactured by splitting water. A
Biofuel novel option, H2-generation from lignocellulosic biomass, based on renewable resources is currently in a pilot-
Hydrogen
scale demonstration or at a commercial stage. The present study reviews the thermochemical, biological, and
Economics
Thermochemical
electrochemical approaches used for biomass-to-hydrogen. The advantages, limitations, and major improve­
Electrochemical ments of each process are presented. A techno-economic assessment is also established based on the production
Fermentation cost, technology readiness level, and industrial scalability.
Biomass The objective is to allow industrial producers to visualise the degree of technological maturity of each option,
Conversion clarify the necessary development efforts before reaching the commercial stage, determine the most relevant and
competitive routes, and assess the suitability of biomass as a feedstock for renewable hydrogen production.
In the reviewed results, the thermochemical process, particularly gasification, partial oxidation, and steam
reforming, presented the best yield for H2 production. Steam gasification is the best compromise because it is
suitable for wet and dry biomass, and it does not require an oxidising agent. As for biological conversion, dark
fermentation is more worthwhile than photo-fermentation due to its lower energy consumption. Additionally,
the electrochemical process is feasible for biomass.
The findings of this study indicate that biomass-hydrogen-based processes are promising options that
contribute to the H2 production capacity but require improvements to produce larger competitive volumes.

1. Introduction applications, mostly ammonia production (27%) and methanol


manufacturing (11%) [11].
Transport (particularly road and air transport) alone is responsible Market forecasts underline that H2 production will increase signifi­
for more than 20% of anthropogenic CO2 emissions [1–3]. To address cantly over the next few years, with a growth of 5–10% per year from 50
this substantial environmental problem, the use of renewable energies, to 82 Mt by 2050 [12]. The industrial H2 demand is also expected to
including alternative fuels and biofuels, should reduce the overall CO2 increase, especially for ammonia and steel production, by about 2 EJ/a
emissions in the transport sector. Forecasts through 2050 underline the for each of these activities in 2050, while the H2 demand for
need for all categories of renewables to account for approximately 27% hydrogen-based vehicles is estimated to reach 22 EJ/a by 2050 [12].
of the total transport fuel consumption [4]. These forecasts also demonstrate an increased use of hydrogen in the
Hydrogen has been closely linked to the transport and fuel sector for manufacturing of new biofuels (from algae or plant materials) that will
many decades. Not only is this molecule involved in several operations impact the hydrogen market and require higher production volumes in
(hydrotreatments) in crude oil refinery but hydrogen is also exploited as the next few years. With a suitable policy and creation of new in­
a reagent/processing agent for the production of alternative fuels (e.g, frastructures, a hydrogen economy could emerge after 2030 [13].
liquefaction of micro-algae) [5–7]. Hydrogen can also be used as a Currently, more than 98% of the H2 production is ensured from fossil
“clean” fuel as it is; its combustion generates only water vapor as a resources, through either natural gas steam reforming, SMR (accounting
by-product [8–10]. Besides crude-oil refining (representing more than for 76% of the global production), or coal gasification (22%) [14]. SMR
33% of the H2 consumption), hydrogen has been utilised in many other is currently the most economical option and is responsible for

* Corresponding author.
E-mail address: a.richel@uliege.be (A. Richel).

https://doi.org/10.1016/j.biombioe.2020.105920
Received 12 March 2020; Received in revised form 24 November 2020; Accepted 3 December 2020
Available online 8 December 2020
0961-9534/© 2020 Elsevier Ltd. All rights reserved.
T. Lepage et al. Biomass and Bioenergy 144 (2021) 105920

approximately 9 kg of CO2 per kg of H2 produced [14]. The alternative CO2 produced from the hydrogen fossil process) [8].
to renewable feedstock instead of natural gas appears to be compulsory Currently, several types of technology underway for the production
to mitigate the release of greenhouse gases and would be a substantial of H2 utilising (lignocellulosic) biomass, such as thermochemical pro­
step toward carbon-neutral emissions [8,15,16]. However, only 2% of cesses, biological conversions, and potentially electrochemically-
the current H2 production is from renewable resources, mostly from assisted production.
water electrolysis, an emerging technology, with no significant use of
vegetal/algal biomass on the industrial scale currently [11]. Water 2.1. Thermochemical productions
splitting is one of the most studied processes, mostly electrolysis (TRL 9,
available at a small scale), with more innovative approaches such as Thermochemical conversion is the most advanced technology for
thermolysis and photolysis (TRL 1–2, under research) [17]. Considering hydrogen production from biomass. The process was established based
biomass, research has emphasised lignocellulosic matrices (e.g. wood, on similar methodologies performed on biofuels, such as biomethane,
forestry side-products, agricultural residues, and dedicated energy adapted from steam methane reforming (SMR) [26]. The three primary
crops) and their conversion, primarily using thermochemical (gasifica­ thermochemical routes are gasification, pyrolysis, and aqueous phase
tion, pyrolysis) pathways [18–20]. The maturity level of these proposals reforming [27].
is aligned with TRL 7. Other alternatives include biological conversion Gasification is a highly endothermic process conducted in an
(TRL 4–5) and electrochemical conversion (TRL 2–4). The hydrogen oxygen-deficient medium at approximately 1000 ◦ C. It consumes an
production costs from biomass remain quite high, ranging from 1.21 to oxidising agent to produce syngas, a synthesised gas composed of
2.42 $/kg when choosing gasification to 1.21–2.19 $/kg for pyrolysis, hydrogen, methane, carbon monoxide, nitrogen, and carbon dioxide.
which is three times higher than that of SMR (0.75 $/kg) [21]. Even if The process differs according to the oxidising agent used and can be
some scenarios involve an economic production of biomass in 2050, designated either as air gasification, oxygen gasification, or steam
biomass should be at the price of coal (ranging from to 40–100 $/MT for gasification [28]. In addition, minor organic and inorganic impurities
coal and up to 140 $/MT for biomass) to be competitive for energy have been found in the syngas. The organic compounds include light
production [22]. Currently, lignocellulosic biomass is expensive to hydrocarbons (LHC), such as CH4, and tar (viscous liquid composed of
produce. However, forecasts indicate a decreasing price of biomass at condensable organic compounds), and the inorganic molecules include
the expense of fossil resources that could be impacted by environmental H2S, HCl, NH3, and alkali metals [29].
policies and the CO2 price [23]. Hence, an increase in the H2 costs from Form an operational point of view, the moist feedstock is introduced
coal gasification is expected from $ 1 to 2.7 $/kg in China by 2030 due to in the gasifier to dry, releasing water (drying zone). Then, the feedstock
taxation [11]. Regarding the cost of hydrogen produced from water by undergoes a pyrolysis reaction (pyrolysis zone), liberating the gaseous
electrolysis, it should decrease from $ 3 to 2.8 $/kg H2 if renewable compounds (CO, CO2, LHC) and tar. Reduction reactions finally occur,
electricity is used and from 5.3 to 4.8 $/kg H2 using the electricity grid and the combustion of the resulting solid from the biomass (called char)
by 2030. These costs would remain higher than those from the biomass produces hydrogen. The unconverted char is transported to the
conversion forecasts [11]. combustor to form CO2 and provide heat for the process. The general air
This review aims to describe the different H2-production pathways gasification reaction of biomass is presented in Equation (1) [21].
from biomass and is formatted as follows: In the second section, the two
Biomass + Air→N2 + CO + H2 + CO2 + CH4 + H2 O + LHC + Tar + Char
important thermochemical and biological pathways are reviewed as
well as the description of the electrochemical conversion potential. A (1)
comparison between the processes in terms of their performance in The “direct route”, based on the direct conversion of the raw
hydrogen yield and their advantages and drawbacks is proposed in the biomass, is an overall simple process suitable for a wide range of
third section. Recent improvements and process optimisations are lignocellulosic materials. However, the relatively low percentage of
reviewed in the fourth section, while a techno-economic assessment hydrogen in the raw biomass (3–11 wt%), as well as the heterogeneity of
based on the latest advances in the field is established in the fifth section. the composition, are responsible for a variable and inefficient yield of H2
An evaluation of hydrogen as a fuel is examined in the sixth section, and and a varied quantity of impurities, such as tar and char [30]. Man­
key findings are highlighted in the seventh section. denoğlu and co-workers revealed the difference between the hydro­
thermal catalytic gasification of lignin and cellulose, the major
2. Main production routes for H2 production from biomass components of biomass. Using K2CO3 as the catalyst at 600 ◦ C, a biomass
conversion to gaseous products of 81.2 wt% can be reached with the
Currently, 98% of the global hydrogen used is produced from fossil cellulose, whereas only 61.4 wt% is obtained from lignin. Hydrogen
fuels, with the predominant production process as steam methane yields of 30.7 mol H2/kg are higher than those of alkali lignin (23.7 mol
reforming. Other options based on renewable resources are currently in H2/kg) [31]. Therefore, biomass containing a high amount of cellulose
pilot-scale demonstrations or at the commercial stage, namely water and a low amount of lignin should favour H2 production.
splitting and H2-generation from biomass, particularly lignocellulosic In conventional gasification using a fixed-bed reactor, air (oxidising
feedstock. agent) is used as a combustible for the partial burning of biomass,
Lignocellulosic biomass is derived from agri-food side products, resulting in a low conversion into H2 due to the high N2 and CO2 con­
agricultural residues, energy crops, marine residues, and forest by- tents in the gas produced. A significant challenge of air gasification is the
products [24]. Dahmen et al. reported that the annual production of removal of nitrogen, a non-combustible gas that reduces the heat value
lignocellulosic agricultural residue has been estimated as approximately of the fuel. As an alternative, gasification can be achieved using oxygen
4.6 billion tons, of which 25% is used, and approximately 7 billion tons or steam instead of air [32]. Oxygen gasification produces a gas of
are produced from grass and forest land, which is used for energy pro­ greater purity, containing a higher level of H2, no nitrogen, and less tar
duction [25]. Biomass has attracted attention for its abundance and low and char. Wang et al. studied an oxidising agent in the gasification of
cost. pine sawdust pellets using a pilot-scale two-stage gasifier. They
Hydrogen production from biomass allows an increase in production demonstrated that an increase in O2 from 20 to 100% induced an in­
quantity, contributing to the economic capacity, increasing source crease in the H2 concentration in the gas produce from 16% to 36% [33].
flexibility, and reducing greenhouse gas emissions. Navaro et al. re­ However, the cost of oxygen makes the process too expensive to
ported that biomass can be carbon neutral due to its maintained natural implement at an industrial scale [34]. Steam gasification is used as a
cycle. The CO2 released during the biomass-derived hydrogen process is compromise between air and oxygen gasification. The H2 content in the
used during photosynthesis for plant growth (without considering the gas is similar to that from oxygen gasification, whereas the process is

2
T. Lepage et al. Biomass and Bioenergy 144 (2021) 105920

cheaper. Nevertheless, tar, char, and a minor amount of N2 are present. Table 2
The general steam gasification reaction of biomass is presented in Reactions identified during steam reforming [40–42].
Equation (2). Reactions Reactants Products

Biomass + Steam→H2 + CO + CO2 + CH4 + HC + Tar Pyrolysis Biomass CO, CO2, CH4, C2H4, H2O, C,
( / ) Primary tar
+ Char ΔH O > 0 kJ mol (2) Tar cracking (pyrolysis) Primary tar CO, CO2, CH4, C2H4, H2,
Secondary tar
A comparison of the different gasification processes is summarised in Secondary CO, CO2, CH4, C2H4, H2, Tertiary
Table 1 [21,35,36]. Steam gasification appears to be the most suitable tar tar
process to produce hydrogen based on the H2 percentage in the gas Tertiary tar C, CO, H2
(40%), higher H2/CO ratio (1.6), and reduced impurities compared with Homogenous gas-gas
reactions
air gasification. Concerning the feedstock, steam gasification is feasible Combustion (oxidation) 2CH4 + O2 2CO + 2H2 (ΔH0 = − 220.0 kJ/
for wet biomass (moisture from 5 to 35 wt%), while air gasification mol)
requires a dry raw material [37]. Dry reforming CH4 + CO2 2CO + 2H2 (ΔH0 = +247.0 kJ/
Steam reforming (SR) is a concomitant purification reaction that C2H4 + 2CO2 mol)
4CO + 2H2 (ΔH0 > 0 kJ/mol)
improves the syngas composition during steam gasification by reducing
Steam reforming CH4 + H2O CO + 3H2 (ΔH0 = +206.0 kJ/mol)
the carbon-to-hydrogen mass ratio (C/H) [38]. The optimisation of C2H4 + 2CO + 4H2 (ΔH0 > 0 kJ/mol)
steam reforming has been performed to reduce the volume of light hy­ 2H2O
drocarbons and tar responsible for the corrosion and blocking of pipes Water Gas Shift CO + H2O CO2 + H2 (ΔH0 = − 40.9 kJ/mol)
due to polymerisation and condensation [39]. Different reactions occur Heterogeneous reactions
Carbon oxidation C + O2 CO2 (ΔH0 = − 393.5 kJ/mol)
during steam gasification, which are presented in Table 2 [40–42]. After Partial oxidation 2C + O2 2CO (ΔH0 = − 246.2 kJ/mol)
the drying step, the pyrolysis reaction occurs, and the biomass is con­ Boudouard reaction C + CO2 2CO (ΔH0 = +159.9 kJ/mol)
verted into a gas rich in CO, CO2, CH4, LHC (C2H4), C, and tar (primary). Water gas reaction C + H2O CO + H2 (ΔH0 = +118.5 kJ/mol)
Primary tar is a complex and condensed mixture of polycyclic aromatic
hydrocarbons and oxygenated molecules, alcohols, phenols, and furans
homogeneous reactions [46]. The char formed during the pyrolysis step
[43]. Under the temperature conditions of steam gasification
is converted into CO and CO2 primarily through slow heterogeneous
(700–1000 ◦ C), homogeneous reactions, such as cracking and steam
reactions during steam gasification, requiring the char to remain in the
reforming, occur and alter the structure of these oxygenated molecules.
reactor to maximise the conversion into gas [47]. Char can also promote
The general tar steam reforming reaction is given in Equation (3) [44].
catalytic activity by supporting the catalysts. Many parameters can be
Cx Hy Oz + (x − z)H2 O → xCO optimised to increase the H2 production yield of gasification. These
( /
+ (0.5 y + x − z)H2 (x ≥ z) ΔH O > 0 kJ mol
)
(3) parameters are highlighted in Table 3 [17,21,48].
At the end of steam gasification, the gas predominantly consists of H2
Steam gasification promotes the steam reforming reaction and in­ and CO2. Therefore, CO2-capture systems (CCS) have been integrated
creases the yield of H2 produced compared with air gasification, rather during the process to increase the concentration of H2. The separation
than promoting a combustion reaction. During the reforming reactions, must be conducted in the same reactor in which the catalysts enhance
the primary tar is cracked into secondary tar, and then into tertiary tar. the WGS reaction to promote the capture of CO2. This process is called
At extremely high temperatures (~1250 ◦ C), it is possible to eliminate sorption-enhanced hydrogen production (SEHP). H2 can be concen­
all the tar [45]. Reforming reactions take place on LHCs in the presence trated up to 99.9% using a pressure swing adsorption system (PSA). By
of H2O and CO2 to form CO and H2. During tar cracking, LHCs (CH4 and varying the pressure, the PSA allows the separation of a gas mixture
C2H4) and carbon oxides (CO and CO2) are generated together with H2. based on the different affinities for each gaseous species to be adsorbed
These compounds are stable under the operating conditions of crack­
ing/reforming. Moreover, the water gas shift (WGS) reaction converts
Table 3
CO in presence of H2O to a mixture of CO2 and H2. Steam added to the
Parameters affecting H2 yield in steam gasification [17,21,48].
gasification process enhances the reversible reaction by shifting the re­
action to H2 production. However, the WGS reaction is exothermic, and Parameters Effects

according to the Le Chatelier’s principle, an increasing temperature will Biomass type High cellulose and lignin content increase gaseous
shift the reaction to the reactants, resulting in the formation of CO and products. High cellulose content biomasses are more
easily gasified with shorter residence periods [49,50].
H2O. Finally, heterogeneous reactions, such as carbon oxidation to form
Biomass particle size Small granulometry provides a larger contact surface and
CO2 or CO, and the Boudouard reaction occur to a lesser extent than a better heat transfer during gasification, improving the
homogeneous reaction (Steam reforming, WGS).
Decreasing granulometry increases H2 yield and decreases
Table 1
tar and char content [51,52].
Comparison of the gasification processes as a function of the oxidising agent (air,
Temperature Increasing the temperature produces a higher H2 yield
oxygen, or steam) [21,35,36]. due to complete homogenous reactions [51,53].
Air gasification Oxygen Steam gasification Increasing the temperature decreases the tar content in
gasification the gas produced [45].
Steam-to-biomass Increasing S/B exhibits a higher H2 yield. Low S/B
a
Products N2, CO, H2, CO, H2, LHC (CH4, H2, CO, CO2, LHC ratio (S/B) promotes CH4 and carbon formation (reformed at higher
CO2, C2H4), CO2 (CH4, C2H4) S/B). Need to be optimised for each process [17,54].
LHC (CH4, Catalysts Increase H2 and CO production by shifting reforming
C2H4), H2O reactions and the WGS reaction [54]. Decrease tar content
Tar (g/kg) 3.7–61.9 2.2–46b 60–95 by promoting the conversion reaction (increase H2 yield).
Average H2 15% 40% 40% The most suitable catalysts for gasification reaction are
composition Ni-based catalysts, alkaline metal oxides, olivine, and
H2/CO ratio 0.75 1 1.6 dolomites [55].
Heating Value 4–7 12–28 10–18 Nickel and cerium catalysts prevent carbon deposition
(MJ/Nm3) and coking [56].
a Residence time Longer residence time increases biomass conversion to
Molecules are listed in decreasing order of presence in the gaseous product.
b gas [57].
Steam-O2 mixture.

3
T. Lepage et al. Biomass and Bioenergy 144 (2021) 105920

on a liquid or solid. Calcium oxide CaO is currently the most commonly given as Equation (7) [69].
used CO2 acceptor (forming CaCO3) and can be easily regenerated [58, ( / )
CHx Oy + (2 − y)H2 O → CO2 + (2 − y + 0.5 x)H2 ΔH O > 0 kJ mol
59]. The method is called “chemical looping” and releases H2 (gasifi­
cation) and CO2 (CaO regeneration) in separate reactors [28]. Scrubbing (7)
is a separation technique used to separate the gaseous impurities, such as
(with​ x ​ = ​ element​ molar ​ ratio​ H/C ​ and​ y ​ = ​ elemental ​ molar​ ratio ​ O/C)
H2S or water [60]. This operation occurs prior to the PSA and is an
energy-consuming separation method [61]. Steam reforming is an
There are two temperature regions in ScWG systems: high temper­
interesting complementary system integrated with existing processes to
ature, typically in the range 500–700 ◦ C and low temperatures of
reduce the tar and char quantities, as summarised in Fig. 1.
374–500 ◦ C. At high temperatures, pyrolysis reactions take place before
Partial oxidation (PO) is an alternative approach to steam reforming
gasification, splitting the feedstock without catalyst contribution.
reactions. The reaction scheme is similar to steam reforming, except that
Nevertheless, alkali catalysts can be added to increase the H2 yield by
H2O is replaced by O2 [19]. One of the advantages is that reactions with
promoting WGS reactions. At low temperatures (374–500 ◦ C), hydro­
oxygen are exothermic, which means there is no need for either an
lysed compounds are gasified and reduced by alkali-based catalysts,
external energy source or the use of more-compact reactors. Different
transition metals, and activated carbon catalysts, which are necessary to
reactions occur depending on the oxygen concentration in the reactor.
enable a high conversion of biomass to gas [70,71]. Moreover, a
PO is widely documented for ethanol, the most described feedstock in
methanation reaction occurs at low temperatures and consumes the
the literature. When hydrogen is produced with carbon monoxide as a
hydrogen produced to form CH4, requiring charcoal or another
by-product from ethanol and low O2 quantities, the reaction is endo­
carbon-based catalyst to solve this issue.
thermic (Equation (4)) and becomes exothermic when the percentage of
Pyrolysis is another tool, besides gasification, for biomass thermo­
O2 within the reactor increases (Equation (5)). In this case, the partial
chemical conversion. Pyrolysis is similar to gasification but can be
oxidation reaction is autothermal, and carbon dioxide is produced. Due
performed at lower temperatures and without an oxidising agent [72].
to the highly exothermic reaction, PO is carried out at lower tempera­
In some specific cases, a small amount of an oxidising agent can be
tures [62,63]. In the case of lower amounts of oxygen, water gas shift
added to facilitate heat generation [73]. Pyrolysis typically occurs at
(WGS) reactions could improve the H2 yield of partial oxidation with
temperatures ranging between 400 and 800 ◦ C, under a pressure of up to
steam (Equation (6)). This reaction is mildly exothermic, and the water
5 bar. According to the operating temperature, pyrolysis can be divided
added to the system avoids deactivation of the catalysts by limiting coke
into three classes: conventional (or slow) pyrolysis, fast pyrolysis, and
formation, as in steam reforming [64,65]. Catalysts are also necessary in
flash pyrolysis. Conventional pyrolysis is carried out at temperatures
partial oxidation, as in steam reforming, except that no external heat
below 450 ◦ C and results in a high charcoal content. Fast pyrolysis
source is required in PO.
produces a high bio-oil yield of up to 75 wt% at medium temperatures
( / )
C2 H5 OH + 0.5 O2 →3H2 + 2CO ΔH O = + 14 kJ mol (4) (450–600 ◦ C) with a high heating rate (approximately 300 ◦ C/min) and
a short residence time [74]. Flash pyrolysis is similar to fast pyrolysis but
( / )
C2 H5 OH + 1.5 O2 →3H2 + 2CO2 ΔH O = − 510 kJ mol (5) at higher temperatures (above 600 ◦ C) and higher heating rates
(>1000 ◦ C/s), while the residence time is shorter (below 1 s), and is used
( / )
C2 H5 OH + 0.5 O2 + 2H2 O→ 5H2 + 2CO2 ΔH O = − 68 kJ mol (6) to maximise the gas yield [42]. However, the fast and flash pyrolysis gas
yields are lower compared with gasification, especially because of the
Supercritical water gasification (ScWG) is an alternative thermo­ lower operating temperatures [75,76]. A general pyrolysis reaction is
chemical route that has been developed at the laboratory scale to be presented in Equation (8) [41].
more robust for the biomass type, including wet biomass (moisture >
35%) such as wood and carbohydrates. Water requires a temperature Biomass + heat→H2 + CO + CO2 + CH4 + H2 O + bio − oil
( / )
above 374 ◦ C and a pressure higher than 221.2 bars to become a su­ + charcoal ΔH O > 0 kJ mol (8)
percritical fluid. Under these conditions, the dielectric constant of water
decreases as well as the quantity of hydrogen bonds. Organic com­ Oxygenated molecules (aldehydes, ketones, alcohols, phenolic
pounds and gases are miscible in supercritical water at high tempera­ compounds and carboxylic acids), water, and ash may be condensed
tures, facilitating their conversion. The total reaction is endothermic and after pyrolysis to form a complex liquid fraction similar to tar called
similar to that involved in aqueous-phase reforming. Residence times “bio-oil”. Bio-oils can be divided into two classes: water-insoluble frac­
can be very low compared with other gasification processes (2–6 s), and tion and water-soluble fraction. The insoluble fraction can be cracked
the reaction can be conducted at a lower temperature (600–650 ◦ C) into platform molecules such as BTX or olefins and can be used in ad­
[66]. However, to maintain these conditions, high energy levels are hesive applications [77]. Concerning the soluble fraction, a steam
needed, increasing the costs and hindering its scalability [8]. Therefore, gasification system can be added to the process to increase the H2 pro­
no system has been developed on an industrial scale because the high duction yield.
pressure required induces substantial capital and operational expendi­ Pyrolysis–steam reforming is a two-stage thermochemical route that
tures [67]. Solving certain technical issues (waste neutralisation, reactor has been extensively studied recently [78]. The catalysts used are the
design standardisation, and comprehension of chemical kinetics) are same as for the cracking steps of gasification and are also inactivated by
compulsory for the development of commercial and viable ScWG sys­ carbon deposition (coke). A catalyst regeneration step enables the con­
tems [68]. The general biomass conversion equation for the ScWG is version of coke into CO2 by combustion, unblocking the active sites [8].
Global H2 yields are similar to those for steam reforming by gasification
and range between 70% and 80% [79]. The separation of the reactors in
the pyrolysis-steam reforming process avoids reforming catalyst inhi­
bition by coke deposition [54]. The process is easy to scale up and offers
an alternative to direct gasification and bio-oil reforming [78].
Aqueous phase reforming (APR) is the third thermochemical
approach used for H2 production from biomass. APR converts mainly
oxygenated compounds into hydrogen. Feedstock molecules are dis­
solved during the aqueous phase and react with water molecules at low
Fig. 1. Schematic illustration of steam gasification of biomass combined with temperatures (<270 ◦ C) and high pressures (up to 50 bar) [80,81]. APR
stream reforming. HC refers to hydrocarbons. processes are suitable for biomass-derived oxygenated hydrocarbons

4
T. Lepage et al. Biomass and Bioenergy 144 (2021) 105920

with a C/O ratio of 1:1 and are miscible in water, such as methanol, CO2) in the dark by oxidising CO and reducing H2O through an enzy­
ethanol, ethylene glycol, glycerol, glucose, or even polyols, such as matic pathway, according to Equation (9) [83].
sorbitol [82]. The small range of effectively soluble molecules converted ( / )
CO + H2 O⇋CO2 + H2 ΔH O < 0 kJ mol (9)
in the process is the principal drawback. However, reforming pure cel­
lulose and woody biomass through APR is possible after a pre-treatment Different microorganisms can perform BWGS. Among the gram-
step [27]. This involves the hydrolysis of cellulose and hemicellulose to negative bacteria, Rhodospirillum rubrum is the most widely explored,
monomers that can be used as feedstock [83]. The reaction mechanism is while Carboxydothermus hydrogenoformans is the most used gram-
similar to steam reforming and relies on C–C bond cleavage to produce positive microorganism [89,90]. An advantage compared with a ther­
H2. The ideal pathway for the conversion of ethylene glycol, whose mochemical WGS reaction is the lower temperatures and pressures
mechanism is well known, is depicted in Fig. 2. In contact with the (30–70 ◦ C depending on the strain, atmospheric pressure). Nevertheless,
surface of a catalyst, the C–C bonds of ethylene glycol are cleaved to different parameters still need to be optimised. Carbon monoxide
produce carbon monoxide. The catalyst also promotes the WGS reaction, released from the feedstock needs to be limited because it decreases
leaving a low amount of CO remaining in the gas stream (<300 ppm). bacterial activity. In addition, microorganisms need to be improved to
Other reaction intermediates are possible, favouring the C–O bond increase the concentration in the bioreactor. Finally, issues concerning
cleavage, leading to alkane (CH4, C2H6) formation and a lower H2 pro­ the mass transfer of gas to liquid should be addressed [90].
duction yield [84]. Dark fermentation, another biological option for producing H2,
Choosing a metal (M)-based catalyst depends strongly on the sta­ occurs when anaerobic microorganisms, such as micro-algae or specific
bility of the M − C bond (enhancing the C–C cleavage pathway) and the bacteria, are sustained in the dark at temperatures between 25 and
M − O bond (enhancing the C–O cleavage pathway). However, the price 80 ◦ C, or even at hyperthermophilic (>80 ◦ C) temperatures, depending
of metal catalysts is prohibitive for large-scale development. Moreover, on the strains. Under these conditions, the gas produced contains H2,
the support of catalysts must be considered. Basic and neutral supports CO2, and small amounts of CH4, CO, and H2S, depending on the con­
are preferred to acidic supports because their selectivity for H2 is higher verted substrate. Hydrogen is primarily produced from the anaerobic
and lower for alkanes, respectively. Indeed, alkanes are produced under metabolism of pyruvates generated during the catabolism of carbohy­
acidic conditions arising from dehydrogenation, and it is critical to opt drates [91]. Compared with photolysis, the H2 yield in the gaseous
for a catalyst that is non-selective to alkanes while working at suitable product is lower due to the minor gases [85]. The yield of hydrogen also
pH conditions [80]. Feedstock biomass and the process conditions also relies on the by-products of the reaction. From glucose, considered as a
influence the selectivity of APR [66,84]. The conversion of complex model substrate, 4 mol of H2 per mol of glucose can be obtained when
oxygenated molecules, such as carbohydrates, is a challenge because acetic acid is the by-product of the reaction and 2 mol of H2 per mol of
they yield lower amounts of H2 [80]. The thermal homogeneous glucose when butyric acid is obtained [92]. More hydrogen can be
decomposition of carbohydrates rapidly creates large amounts of coke produced when the reduced end-product is acetate, and that pathway
responsible for catalyst deactivation and competing with the reforming should be promoted over butyrate. However, both molecules still
reactions. contain hydrogen that could be released to increase the yield. It is
possible to direct a specific pathway for H2 production by changing the
2.2. Biological conversion conditions of the bioreactor [85]. Other parameters also influence H2
production, such as the pH, hydraulic retention time (HRT), and gas
Biological conversion can be divided into three categories: biological partial pressure. Another advantage of dark fermentation is the possi­
water gas shift reaction, dark fermentation, and photo-fermentation. bility of converting cellulosic and lignocellulosic materials, such as
Each process depends on the nature of the enzymes used to catalyse wastes, to hydrogen [93,94]. This biological process has the potential to
H2 formation [85]. Commonly, microorganisms, such as micro-algae integrate waste management into energy production. However, it still
and cyanobacteria, produce the enzymes required to synthesise H2, requires the development of current technologies to be financially
including nitrogenases and hydrogenases [86]. Nitrogenases reduce competitive and have the potential for practical application and com­
protons (H+) from adenosine triphosphate (ATP) and electrons and mercialisation. Globally, two factors are limiting the development and
release H2, while hydrogenases reversibly catalyse the conversion of scaling of biological processes. Compared with thermochemical pro­
protons to hydrogen. cesses, the H2 yield and production rate are lower (up to several months)
In contrast to thermochemical processes, biological conversion takes making biological processes less competitive, even if improvement has
place at lower temperatures (30–60 ◦ C) and pressures (1 atm), been made with dark fermentation [95]. In addition, the complexity of
decreasing the energy cost. Moreover, the microorganisms used can be lignocellulosic material and agricultural waste requires the material to
easily regenerated by replication, decreasing the turnover frequency be enzymatically fermentable. This additional time-consuming and
compared with chemical catalysts easily deactivated during thermo­ expensive process hinders the development at an industrial scale. The
chemical conversions [87]. Biological processes are particularly inter­ variability of the feedstock notably affects the H2 yield and the economy
esting in waste management and enable the conversion of agricultural of the process, demonstrating that the selection of substrates is crucial.
waste, agri-food effluents, and solid residues, or municipal solid waste. Photo-fermentation (PF) is the latest biological process used for H2
Investigations of sewage sludge have increased in the last ten years production. PF is catalysed by nitrogenases in purple non-sulphur bac­
because the nature of the material is favourable to its conversion teria to convert organic acids or biomass into hydrogen from solar en­
through different biological processes [88]. ergy in a nitrogen-deficient medium. Equation (10) presents the global
The biological water gas shift (BWGS) reaction depends on the reaction performed on glucose as the model substrate [96].
capacity of photoheterotrophic bacteria, using carbon monoxide as the
C6 H12 O6 + 12H2 O→ + 12H2 + 6CO2 (10)
carbon source. These microorganisms can produce H2 (together with
PF can be performed either through single-step or two-stage
fermentation. The single-stage process is a low-cost approach but suf­
fers from several drawbacks, such as the high energy demand from ni­
trogenases and their low solar energy conversion. The two-stage
fermentation is a system in which dark-fermentation occurs first, and the
organic acids obtained undergo PF sequentially to produce more H2. The
Fig. 2. Mechanism for APR converting a diol component into hydrogen (and disadvantages are the difficult operations to control the different
combined with a WGS post-process).

5
T. Lepage et al. Biomass and Bioenergy 144 (2021) 105920

bacteria and the parameters between the separate stages [97,98]. Recent [105].
studies have investigated PF as waste-prevention process to produce H2 Classic PEMEC systems require an expensive catalyst made of noble
from industrial and agricultural wastes [99,100]. However, it remains metals, that are extremely sensitive to CO and other impurities released
the least financially competitive process among all the biological by biomass during electrolysis. Moreover, PEMECs consume a high
methods [101]. volume of alkaline molecules to neutralise the organic acids produced
during the electrolysis of biomass, thus making the process even more
expensive. The same findings have been reported for MEC systems
2.3. Electrochemical conversion possibility where microorganisms cannot directly use the biomass. Biomass must be
fermented and converted into organic acids or alcohols, and only those
Electrolysis is an electrochemical process widely investigated for molecules can be converted to H2 through the MEC. In addition,
hydrogen production by splitting water molecules. The mechanism oc­ although its electrical consumption is lower than that of water elec­
curs in a fuel cell (containing a cathode and an anode) at a low tem­ trolysis, the conversion rate of MEC is too slow for the process to be
perature and relies on the flow of an electric current through a competitive. These drawbacks have slowed down the development of
conductive electrolyte (alkali or polymer) in water. This results in the electrolysis systems producing H2 from biomass [106].
splitting of water into O2 and H2 [102]. The conversion is fast, Even if the electrical consumption can be lowered by using bio­
straightforward, and produces pure H2 after separation. Moreover, there ethanol instead of water (e.g. Chen et al. reduced it by 26.5 kWh/kg H2
are no carbon, sulphur, or nitrogen side-product compounds, saving converting ethanol using palladium nanoparticles deposited on titanium
purification costs when compared with those of gasification processes. nanotubes), the total energy balance is impacted by bioethanol pro­
The limitation of water splitting relies on thermodynamic parameters duction. Indeed, this issue decreases the interest in this process because
because an electric consumption of a minimum of 45 kWh/kg H2 is an economically viable bioethanol is challenging to produce from
required to reach a sufficient yield of hydrogen. biomass, especially from lignocellulose [107]. Investigations into the
Electrochemical conversion is also possible for biomass. The differ­ direct electrolysis of biomass should be performed in the future.
ence between water and biomass electrolysis lies in the reaction
occurring at the anode. The feedstock is oxidised instead of producing 3. SWOT analysis of the main process used from biomass
gaseous oxygen from the water. Biomass electrolysis can be achieved
through two different technologies, Proton Exchange Membrane Elec­ Several (thermochemical, biological, and electrochemical) technol­
trolysis Cell (PEMEC) and Microbial Electrolysis Cell (MEC). Both ogies are described below regarding their ability to produce hydrogen
PEMECs and MECs are commonly used for bio-based molecules such as from biomass. Table 4 summarises the key advantages and drawbacks of
ethanol and glycerol. Polymeric molecules, such as cellulose or wood each process.
sawdust, cannot be converted directly by electrolysis. These systems Thermochemical conversions have the advantage of being available
require a single or two chambers containing an anode/cathode couple. at a large production scale because the technologies used are based on
The conversion of organic matter occurs at the anode by an oxidation current well-established methods for converting fossil fuels. Therefore,
reaction releasing protons (H+). A reduction reaction occurs at the the industrial design has already been established [116]. Despite being
cathode, allowing the formation of H2 [103]. PEMEC systems contain a carbon-neutral, thermochemical processes release CO2 and require a
polymeric membrane that transfers protons to obtain a high hydrogen combination with carbon capture systems (CCS) to achieve negative
gaseous fraction [104]. In MEC systems, the oxidation of organic matter emissions. Even if this solution is associated with an increased invest­
to produce H+ is performed using electrochemically active microor­ ment, it is the best economical solution to reach the environmental
ganisms as catalysts, making the process a “bio-electrochemical” system

Table 4
Comparison of the main advantages and drawbacks, and H2 yields (expressed in g/kg feedstock) of several process options.
Processes Advantages Limitations H2 Yield (g/kg Ref.
feedstock)

Thermochemical processes
Biomass Existing industrial design, forest residue and industrial CO2 emissions, tar and char formation leading to catalyst 40–190 [19,75,108,
Gasification waste recycling, high biomass conversion efficiency, no deactivation, H2 variation due to biomass complexity and 109]
expensive oxygen source required in steam gasification composition variations, high operating temperature, need for
catalysts regeneration, high reactor cost
Steam Existing industrial design, no expensive oxygen source CO2 emissions, high operating temperature, need for catalysts 40–130 [62,109]
Reforming required regeneration
Partial Existing industrial design, no catalyst requirement, low CO2 emissions, high operating temperature, high cost of 16–140 [109–111]
Oxidation desulfurization requirement oxygen, adapted only for few molecules
ScWG Suitable on biomass with high moisture content CO2 emissions, high pressure needed, high energy consumption 20–40 [112]
APR Suitable for feedstock in aqueous solutions CO2 emissions, adapted only for few molecules, need for pre- 1–40 [113]
treatment
Biomass Existing industrial design, forest residue and industrial CO2 emissions, tar and char formation, H2 variation due to 25–65 [114]
Pyrolysis waste recycling, versatile conversion of biomass (gas, bio- biomass complexity and composition variation, need for
oil, biochar), simple process catalysts regeneration, high reactor cost
Biological processes
Dark Organic and biological waste streams recycling, conversion Low H2 yield and rate production, high by-products generation, 4–44 [77,109,
fermentation of algal biomass (high growth rate), low operating need for pre-treatment 115,116]
temperature and pressure, various suitable carbon sources
Photo- Organic and biological waste streams recycling, nearly Low H2 yield and rate production, high surface area 9–49 [77,97,109,
fermentation complete substrate conversion, low operating temperature requirement, need to control the bacteria, high energy demand 117]
and pressure for enzymes, low energy solar conversion efficiency
Electrochemical processes
MEC No purification required, lower electrical consumption Expensive catalyst, low production rate, suitable for less 15–98 [103,106,
(compared with water electrolysis), mild conditions complex molecules in solution 118]
PEMEC No purification required, lower electrical consumption Expensive catalyst, low production rate, suitable for less / [106]
(compared with water electrolysis), mild conditions complex molecules in solution

6
T. Lepage et al. Biomass and Bioenergy 144 (2021) 105920

objectives [119]. These processes rely on essential catalysts that are (Ficus virens leaves) using limestone/dolomite (1:1) as the catalyst
prone to deactivation. They require high temperatures compared with [126]. Wei et al. highlighted the variations in the H2 yield due to
biological and electrochemical routes. Overall, the best yields are ob­ biomass type and catalyst. They obtained a 10% and 15% increase in H2
tained through gasification. Gasification and pyrolysis are adapted for a concentrations in the gas produced from legume straw and sawdust
wide variety of dried feedstocks, while steam gasification and steam using dolomite compared with inert sand. This increase is attributed to
reforming are the best compromise for wet complex biomass treatment. the catalytic steam reforming of tar. Meanwhile, varying the tempera­
However, the processes are more sensitive to catalyst deactivation by tar ture demonstrated a significant effect on tar yield (g/Nm3) and
and char formation. Partial oxidation and APR are only applicable to hydrogen yield (g/kg fuel) [127]. A significant effect on hydrogen yield
simple molecules in solution, requiring an eventual pre-treatment step was demonstrated by Ma and co-workers in 2019 by studying dolomite
that increases the cost of the processes. The ScWG system requires high and olivine as catalysts at different temperatures. When the temperature
pressures, which also represents a significant cost that hinders scaling up was increased from 700 to 900 ◦ C, the tar yield decreased from 12.5 to
[120]. PO, APR, and ScWG would have less impact because their 7.2 g/Nm3 for dolomite and from 15.9 to 9.1 g/Nm3 for olivine. The tar
applicability at a large scale is limited. conversion led to an increased hydrogen yield from 36.2 to 46.4 g/kg
Biological conversion suffers from low yields and a slow conversion and from 32.4 to 42.3 g/kg of biomass, respectively [128]. The choice of
rate. The production from complex biomass requires pre-treatment for gasifier depends on the biomass feedstock and the conditions for H2
prior fermentation, hindering the processes at the pilot scale. They are production. Therefore, development and optimisation steps are required
more suitable for rich liquid streams, such as molasses or food waste for each raw material.
[121]. Moreover, compared with APR and ScWG, the conditions of the Currently, the best compromise consists of gasification using steam
fermentative processes are milder, and the hydrogen yields are more as an oxidising agent in terms of gas composition, impurities, and cost.
substantial due to a higher efficiency, especially in the case of dark Recent developments in steam gasification have focused on decreasing
fermentation (80%). Even though PF has a slightly higher yield than the temperature processes and the tar and char formation through
dark fermentation, the lower efficiency (0.1%) decreases the interest in catalyst use to improve the H2 yield [42]. Moreover, steam gasification
this technology. In addition, dark fermentation is more worthwhile allows more flexibility regarding the feedstock, allowing the use of wet
considering its higher production rate and lower energy consumption. biomass without requiring expensive oxygen. The primary objectives are
Therefore, dark fermentation or a combined process with PF should be the development of improved catalysts, especially dolomite and olivine,
favoured. to reduce the operating temperature (the total energy used), tar content,
Concerning electrochemical conversion, water electrolysis does not and therefore, the cost of the process.
directly cause carbon emissions but has a high electrical consumption Linked to gasification improvement, most of the research in terms of
(4.5-50.6 kWh/m3 H2) [122]. Biomass electrolysis has been investigated steam reforming (SR) is focused on new catalyst discovery, temperature
because lower electrical consumptions have been obtained with the optimisation, and char and coke degradation. Guo et al. increased the
development of the MEC and PEMEC systems (0.6–1 kWh/m3 H2) [123]. yield by using rice husk char (RHC) as a support metal catalyst, from
This consumption is even lower than the energy required for biological 196.6 to 269.6 mL/g using K-RHC, 274.9 mL/g for Cu-RHC, and 342.7
processes, while having similar conversion rates. Although MEC and mL/g for Fe-RHC. Char and char-supported catalysts promote H2, CO,
PEMEC are interesting methods for waste management, the low and CH4 production [129]. In addition, a high C/H ratio of biomass
hydrogen production rate is the most significant impediment to these induces coke formation during gasification; a carbon deposit negatively
methods being competitive [124]. affects the catalyst. Decreasing the C/H ratio is driven by steam
To supply the necessary amount of biomass to produce hydrogen, reforming, thereby reducing coke formation. In 2018, Yaghoubi et al.
two types of feedstock can be used: lignocellulosic residue and dedicated studied the effect of different parameters on hydrogen production from
crops. Lignocellulosic waste is available at a low cost because they are biomass in a dual fluidised bed gasifier. They demonstrated that H2
generated at the end of the harvest season for such crops as cereal wheat production could be optimised at operating temperatures between 800
or during the transformation process. However, dedicated energy from and 820 ◦ C. At higher temperatures (>820 ◦ C), hydrogen production
crops, such as sorghum, requires land use and growth time, which can be decreases due to char combustion for energy input; thus, char is not
a challenge. As mentioned by Kammoun et al. it is possible to group converted into H2. Residence time influences the specific gas formed
biomass types according to their chemical composition (carbohydrates, during gasification, which is directly linked to the steam/biomass ratio
lignin, and others) to facilitate the pre-treatment required [24]. (S/B). Typically, an S/B ratio of 1.3 allows the optimal production of H2.
Thus, biomass represents a renewable precursor to hydrogen pro­ Higher S/B ratios decrease the residence time and do not allow complete
duction and could be comparable to water electrolysis as an alternative gasification reactions, thereby producing less H2 and more
to fossil feedstock for chemical production and even in the future for fuel carbon-containing gases. Finally, decreasing the reactor diameter and
transportation. biomass particle size increases the H2 yield by gasification [130]. The
general concentration of H2 in the reported gas produced varies between
4. Recent developments and perspectives for existing processes 18% and 59%. This variation is due to the different catalysts used, and
studies have focused on their efficiency. Ni-based and char alkali earth
4.1. Improvements of thermochemical approaches metals are the most effective and explored catalysts. For example, Liu
et al. studied catalytic tar cracking from rice hull gasification and ob­
Gasification is perhaps the most thoroughly studied process, and its tained a tar conversion of 94.4% with a Ni-based catalyst (Ni6/PG),
improvement has been abundantly described and updated. One key enabling a hydrogen yield of 57.7% [131]. Al-Rahbi et al. used tire
element of the current research is to find suitable operating conditions pyrolysis char as a catalyst for the gasification of wood pellets. The H2
and catalysts that can effectively increase the volume of the gas pro­ concentration increased up to 50% and the H2 yield from 8.4 to 39.2
duced and its quality at a low cost [125]. New efficient catalysts are mmol/g biomass when the temperature was raised from 700 to 900 ◦ C
added to promote tar conversion and prevent the formation of unwanted [132]. Yao et al. demonstrated that biochar is a promising catalyst and
products. Some studies have already reviewed and compared a wide support for biomass gasification. Biochar from CSP allows a good
range of available catalysts for thermochemical processes, especially interaction with the volatiles due to the alkali and alkaline earth metals
gasification [21]. Usually, these are dolomite, olivine, potassium min­ present, facilitating the conversion to H2. The addition of Ni (Ni-based
eral, and nickel-based compounds [108]. Recent results are highlighted biochar) improved the production of hydrogen from 45.91 to 92.08 mg
for these catalysts. According to Zhang et al. for a range temperature H2/g biomass at 800 ◦ C in a two-stage fixed-bed reactor during steam
from 450 ◦ C to 850 ◦ C, the maximum H2 yield was 204.6 mL/kg biomass reforming [133]. Xu et al. also demonstrated the effect of alkaline earth

7
T. Lepage et al. Biomass and Bioenergy 144 (2021) 105920

metals on gasification efficiency. Calcined calcium oxide (CaO), gaseous products. Moreover, catalysts also improve the yield of H2
comprising 5 wt% of Fe gave the highest H2 yield (26.40 mol/kg production, tar decomposition, and biomass conversion. Carbonates
biomass) in a fluidised bed reactor at 700 ◦ C for wood sawdust gasifi­ (Na2CO3 and CaCO3) and metal oxides (Al2O3 and Cr2O3) have been
cation. The Fe/CaO catalyst is less efficient for gasification than some studied and compared for different biomass species. The most effective
other catalysts and for CO2 capture compared with CaO. However, Fe catalyst also depends on the biomass type selected as the feedstock
impregnation prevents CaO deactivation and allows for a better tar [142]. Research has also focused on the discovery of new Ni-based
conversion, giving a higher H2 yield compared with CaO only (12.36 catalysts and optimal operating conditions according to each type of
mol/kg biomass) [28]. Overall, the key perspectives in terms of R&D biomass. Akubo et al. studied the conversion of six biomass wastes (rice
reside in the discovery of cheap and bio-based catalysts that resist coke husks, coconut shells, sugarcane bagasse, palm kernel shells, cotton
deactivation while being easily regenerated. This could allow a better stalks, and wheat straw) and three biomass model components (cellu­
gas quality (H2 proportion), lower tar and char content, and a process lose, lignin, and xylan) into hydrogen by pyrolysis-steam reforming. In a
requiring a lower amount of energy. two-stage fixed-bed reactor system, pyrolysis occurs at 550 ◦ C and steam
Regarding partial oxidation (PO), several studies have been con­ reforming at 750 ◦ C with a NiAl2O3 catalyst. The catalyst and steam
ducted to improve partial oxidation using noble metal catalysts, such as enable a H2 yield of 25.35 mmol/g for palm kernel shell conversion,
nickel- and cerium-based catalysts; however, their cost remains a which represents 57.36 vol% of the gas composition. For the biomass
limiting factor [61,134]. Ma and Muller investigated various catalysts components, lignin exhibited the highest H2 yield (25.25 mmol/g)
for the partial oxidation of tar from biomass gasification. They revealed compared with cellulose (19.72 mmol/g) and xylan (20.54 mmol/g).
that NiO, a nickel-based catalyst, was more active in the conversion of Different mixtures of lignin/cellulose/xylan led to the conclusion that
naphthalene, a tar model compound. A conversion yield of 95% was the more the feedstock is lignified, the more hydrogen is produced,
obtained at 600 ◦ C in a fixed-bed reactor [135]. Using Rh as a catalyst while cellulose and hemicellulose pyrolysis release more CO and CO2
(0.5 wt%), Kim et al. obtained an H2 production yield of 90% at 750 ◦ C [143]. Chen et al. tested Ni/CaAlOx as catalysts on wood sawdust
from dimethyl ether partial oxidation in a microreactor. The study through pyrolysis (T = 500 ◦ C) and steam reforming (T = 800 ◦ C). The
highlights the efficiency of their developed microreactor compared with hydrogen yield was 15.57 mmol/g feedstock with a Ca/Al molar ratio of
a conventionally packed reactor [136]. Partial oxidation was also per­ 1:2. Ca is a cheap alkaline metal that improves the conversion of biomass
formed by Zukowski and Berkowicz on methanol in a fluidised bed of a by increasing the CO selectivity, which can be used to control the H2/CO
catalyst containing iron and chromium oxides [137]. Using an equi­ ratio in the syngas. With a Ca/Al ratio of 3:1, a total concentration of H2
molar ratio of CH3OH and N2O, a yield of 95% was achieved at 350 ◦ C. and CO of 90 vol% was reached. However, catalyst deactivation
This method occurs at a lower temperature compared with other ther­ resulting from coke deposition is the limiting factor [144]. Nevertheless,
mochemical processes. However, the reaction co-generates CO2 and N2. according to Jin et al. Ca used in the catalyst also increases hydrogen
Separation steps are necessary following the partial oxidation. Many production. Under the CaO form, CO2 is absorbed during the conversion
efforts have been made to develop and upgrade partial oxidation pro­ process. When Ca was incorporated into the Ni–Mg–Al catalyst, the
cesses at the pilot scale. Compared with steam reforming, PO has a lower hydrogen yield increased from 10.4 to 18.2 mmol/g of wood sawdust
or equal efficiency in terms of H2 production depending on the substrate. [145]. Dong et al. searched for a catalyst promoting hydrogen produc­
For ethanol reforming, yields are up to 50% higher for SR [138]. tion while minimising catalyst deactivation. The NiO–ZnO–Al2O3 cata­
Moreover, PO is more adapted to the indirect route by using lyst obtained by precipitation has been tested with different Ni contents.
derived-biomass molecules with a lower C/H ratio, such as ethanol The H2 yield increased from 8.2 to 20.1 mmol/g wood sawdust when the
(having a similar C/H ratio as methane), and not on lignocellulosic Ni molar ratio was increased from 5 to 35%. Moreover, when the Ni
biomass feedstock. PO can be attractive in order to valorise co-products content was increased to 25 and 35%, coke deposition on the catalyst
from already settled industries, including glycerol from biodiesel pro­ was below 1 wt%. This study demonstrates the activity of Ni for the
duction and ethanol from breweries, producing it as a side product of conversion of biomass and its resistance to deactivation due to carbon
non-alcoholic beers due to an increasing market for this type of product. deposition [146].
For ScWG, current research is focusing on more efficient catalysts. Li Regarding aqueous phase reforming (APR), extensive research is
and Guo evaluated the performance of the Ni/MgAl2O4–Al2O3 catalyst currently being conducted to improve catalyst stability and activity to
in 2018. They also studied the effects of different parameters on the reach commercial applications and scale up. Metal catalysts (M), such as
ScWG, such as the heating rate, temperature, and catalyst loading. An Pt and Pd, are currently the most frequently used, as well as Ni-based
increased H2 production yield was obtained when the temperature was catalysts, because their activity for C–C and C–H bond cleavage is
increased from 200 to 600 ◦ C and reached a yield of 45 mol/kg for a 5 wt high, leading to high H2 production. These catalysts are also necessary
% glucose solution at 600 ◦ C and a residence time of 300 s. CO2 was the for their high activity in WGS reactions. They increase the H2 yield
other significant gaseous product, and a separation process after ScWG compared with Rh and Ru catalysts through a higher propensity for
could deliver a high-purity hydrogen gas (>80%). Temperatures above alkane formation by promoting the C–O bond cleavage pathway. The
500 ◦ C are favourable for H2 production, while CH4 is the predominant thermal conditions and high pressure increase the activity of the WGS
gas produced at temperatures ranging from 400 to 500 ◦ C [139]. Nanda reaction and improve the conversion of oxygenates into H2. Moreover,
et al. compared the activity of different alkali catalysts for timothy grass. thermal decomposition of oxygenates is unlikely to occur, leading to less
The highest hydrogen yield was 5.15 mol/kg of feedstock at 650 ◦ C for coke deposition and catalyst deactivation. H2 can be concentrated up to
45 min with a 1:8 biomass-to-water mass ratio. They showed that an 99.9% using a pressure swing adsorption system (PSA) [80]. Coke
increase in temperature improves the WGS reaction, resulting in higher deposition can be mitigated by converting polyols instead of carbohy­
H2 and CO2 yields. For the WGS reaction also being catalysed by alkali drates. Their resistance to thermal decomposition (Maillard reaction)
catalysts, KOH produced the best results (H2 yield = 8.91 mol/kg), due to the absence of carbonyl enables the decrease in coke deposition
followed by K2CO3, NaOH, and Na2CO3. Alkali catalysts are worth and increases the reforming reactions producing H2. In addition, by
exploring further for the production of H2-rich syngas [140]. increasing the steam-to-biomass ratio, the feedstock is diluted, enabling
Many developments have also been proposed for pyrolysis, in a yield comparable to that of polyols [84]. A solution is the reduction of
particular aiming at elucidating the optimal process conditions. Waheed glucose to sorbitol, which has a better resistance to thermal decompo­
et al. increased the temperature from 850 ◦ C to 1050 ◦ C, inducing a sition (responsible for catalyst deactivation by coke accelerated forma­
hydrogen yield increase from 20.03 to 30.62 mmol/g feedstock [141]. tion). Moreover, metal catalysts, such as platinum-based catalysts, have
Biomass type, particle size, heating rate, pressure, and residence time a higher selectivity for sorbitol than sugars in H2 production because of
can be adapted to maximise the production of charcoal, bio-oil, or the reduction of their aldehyde function to alcohol [83].

8
T. Lepage et al. Biomass and Bioenergy 144 (2021) 105920

4.2. Improvements of biological processes predominantly in the development of a better-performing


photobioreactor.
When considering dark fermentation, the impact of the pre-
treatment is significant. Indeed, it improves the H2 yield by hydro­ 4.3. Recent developments in electrochemical approaches
lysing the feedstock into small-weight molecules that are easily metab­
olised by microorganisms. In 2018, Kumar et al. reviewed the hydrogen Current research is motivated to investigate the use of native biomass
production from various biomass types (lignocellulosic, algae, and as feedstock without the conversion into an intermediate. In 2016, Liu
wastewater) by dark fermentation, while Bundhoo focused only on crop et al. proposed an alternative by developing a new Chemical-Electrolysis
residues [93,147]. For lignocellulosic biomass, different pre-treatments Conversion (CEC) using aqueous polyoxometalates (POM) as catalysts
have been tested as dilute acids. Alkali pre-treatment was carried out to [106]. In this case, POM is in solution and is not coated on the anode.
remove lignin, followed by enzymatic hydrolysis (cellulase and xyla­ POM receives electrons from biomass oxidation and is enhanced by
nase) to obtain fermentable sugars. The combination of alkali and acid heating or light irradiation. As POM has a low standard redox potential,
pre-treatments has also been investigated. Among all the results, the it can be easily re-oxidised, and protons are released and then reduced at
highest yield obtained was 176 mL H2/g cornstalk using cellulase in the cathode to form H2. POM acts as an electron-coupled-proton carrier.
lightly acidic conditions (pH 4.8, 0.2% HCl). Recent studies have POM has the advantage of being a low-cost self-healing catalyst (100,
focused on hydrogen production using microalgal biomass. Lunprom 000 turnovers) and is tolerant to impurities produced from the biomass.
et al. produced H2 through anaerobic solid-state fermentation followed CEC can be achieved in a PEMEC system without noble metal catalysts.
by dark fermentation using the microalgae Chlorella sp. The H2 yield was The energy consumption is extremely low (0.69 kWh/Nm3 H2)
16.2 mL H2/g volatile solid (VS, total organic matter in the biomass) compared with that of water electrolysis (4.13 kWh/Nm3 H2). This
[148]. With the help of pre-hydrolysis with simultaneous saccharifica­ process allows for the conversion of native biomass, such as wood dust,
tion and fermentation, Giang et al. improved the H2 yield up to 172 mL to produce H2 at a low temperature and with low energy consumption.
H2/g VS and productivity (2.4 ml/g VS. h) [149]. Wang and Yin also Another method based on Fe3+ was developed by Yang et al., in 2017
showed the production of hydrogen from different microalgal biomass, using a similar approach to POM [156]. POM has a large molecular
and the effect of the pre-treatment increased the H2 yield by disrupting weight, inducing a low rate of electrochemical conversion. In order to
the microalgal cells (up to 958 mL H2/g VS from Chlorella sp. with improve the conversion rate, they investigated catalysts with smaller
HCl-heat pre-treatment). However, the processes are far from the in­ compounds, such as Fe3+. They reported a novel method to produce H2
dustrial scale, and more studies are required to improve the profitability by electrolysis from biomass using iron ions as the catalyst in the liquid
of such processes [150]. More research is necessary regarding phase. The production of hydrogen was achieved using PEMECs from
pre-treatment to standardise the operating conditions and improve the different biomasses comprising glucose, starch, and cellulose, without
efficiency of the biological process. As reviewed by Zabed et al. different any noble metals used as catalysts. During the process, Fe3+ in the so­
types of pre-treatment exist (biological, chemical, physical, combined lution reacts with the biomass. The oxidised biomass is degraded to low
treatment) [151]. Future work will focus on scaling up to generate molecular weight derivatives, and Fe3+ is reduced to Fe2+. On the
hydrogen in larger quantities from biomass to reach an industrial scale. cathode, Fe2+ is oxidised to Fe3+, and H+ is reduced to H2 on the anode.
Moreover, practical and economic feasibility analyses at larger scales are The Fe ions act as an electron carrier, and their low redox potential al­
necessary to evaluate the applicability of the process [148]. Further lows for low electric energy consumption. Another example of low
investigation is needed into the development and design of a electric energy consumption was investigated by Hibino et al., in 2018.
better-performing bioreactor to improve the conversion. The issue of They produced hydrogen by direct electrolysis of bread, sawdust, and
reducing the cost of the pre-treatment must guide researchers to affect rice chaff at 150 ◦ C using their own anode developed with an
the cost of the process and to set up dark fermentation on a larger scale oxygen-functionalising carbon without any noble metals. The perfor­
[152]. Other research perspectives can be adopted for the identification mance in terms of catalytic activity was similar to that of a conventional
of new strains of more suitable bacteria or microorganisms, revealing Pt/C anode. H3PO4, is added to the medium, hydrolysing the waste
new opportunities [153]. Finally, at first sight, the use of biological biomass and enhancing electrolysis by increasing the solubilisation of
processes in general makes more sense to valorise effluents containing the compounds. The average yield was 0.25 mg H2/mg feedstock for
sugars, such as molasses and vinasse, and wastewater, including sewage each biomass. In terms of perspectives, investigations should focus on
sludge, that can be processed by microorganisms. Research has working with higher biomass quantities to scale-up the electrochemical
increased in the last ten years because the nature of the material is process [157]. Concerning the environmental impact of the process, CO2
favourable to its conversion by different biological processes. The high is released during the production of hydrogen from biomass, contrary to
water content is suitable for biological conversion, but is a drawback for water electrolysis considered alone. However, the net balance for
major processes such as gasification. greenhouse gas emissions is zero when considering the CO2 consump­
For photo-fermentation (PF), different parameters (substrate con­ tion during biomass photosynthesis. In addition, a CO2 capture system,
centration, C/N ratio, and phosphate concentration) were optimised by such as CaO, can be added to improve the H2 concentrations in the
Wang et al. to develop a single-stage process to produce hydrogen by PF produced gas, limit emissions, and increase the total process profit­
from straw biomass using Rhodobacter sphaeroides. The biomass was first ability [59]. Whether the process is applied to water or biomass, the CO2
hydrolysed under acidic conditions using 5% HCl at 118 ◦ C for 30 min. released to produce the electricity used during the process must be
The conversion of the hydrolysate to H2 was achieved in a neutral me­ considered.
dium by two Rhodobacter sphaeroides strains: HYO1 (wild type) and
WHO4 (mutant). They obtained a higher yield (4.62 mol H2/mol 5. Techno-economic comparison of hydrogen producing
reducing sugar for the WHO4 strain) than direct sugar PF due to pH processes
stabilisation [154]. To improve the efficiency of PF, some researchers
have shown that combining dark fermentation with PF can improve The techno-economic assessment of an industrial hydrogen produc­
efficiency of the biological process. Integrated dark- and tion facility relies on two major expenses: capital investments (CAPEX)
photo-fermentation can solve the issues of the individual systems and and operational costs (OPEX). The critical issue is to reduce the CAPEX
give higher hydrogen yields than PF under the same conditions [155]. and OPEX of the different processes while increasing the production
The key perspective concerning PF consists of its combination with dark volume. This approach would allow the reduction of H2 production costs
fermentation. The dual system must be optimised to be implemented at from renewable resources to compete with fossil fuels. In most cases, the
the pilot scale. The major improvement necessary in PF itself resides CAPEX value was estimated using the software “Aspen Plus” or “Aspen

9
T. Lepage et al. Biomass and Bioenergy 144 (2021) 105920

Hysys” in order to simulate the processes and to make economic as­ both the yield and energy consumption. Regarding the oxidising agent in
sessments [158]. Plant size and capacity have a major impact on CAPEX. gasification, steam is more beneficial than oxygen for economical
The larger the facilities, the higher the CAPEX, but leading to a decrease (representing a difference of 7–22 M€ for the CAPEX) and practical
in the production cost [159]. The presence of an integrated carbon reasons (WGS and CO2 removal) [163].
sequestration or carbon capture and storage (CCS) system also signifi­ Among the other thermochemical processes, ScWG and APR have
cantly affects the total production cost, increasing the CAPEX due to the been developed up to the pilot scale (TRL = 4–5). Concerning the ScWG,
supplementary technology used. The advancement level of the tech­ the conversion of wet biomass, such as algae, sewage sludge, or organic
nology or technology readiness level (TRL) and the availability of wastes, avoids the energy-intensive drying step. However, the key issue
structures for the conversion processes are also critical factors to discuss is the high pressure required for the process. The resulting high energy
[109]. Concerning the OPEX, the price of the biomass feedstock, which consumption and the specific installations increase both the OPEX and
is mostly related to its availability and transport, has a significant in­ CAPEX, making the process less competitive. Özdenkçi et al. determined
fluence on the production cost. Other parameters inherent to biomass that the OPEX is an annual cost (60 M€) equivalent to the total instal­
need to be considered, including the type of feedstock and the eventual lation cost for the ScWG of black liquor. Based on the simulation, the
need for the pre-treatment step. An ideal biomass must be as cheap as current H2 production cost by SCWG is primarily affected by the OPEX
possible, resulting in a sustainable yield upon its use. In addition, the and could be decreased by developing large installations, such as for
process conditions have a significant impact on the cost of the process. gasification [168]. APR is performed on simple molecules, such as
The cost includes the energy delivered to increase the temperature and ethanol. Its sustainability relies on the cost of reliably produced bio­
pressure, the electricity used for equipment or reactions (electrolysis), ethanol. Currently, alcohol production from lignocellulosic biomass is
and the type and cost of the catalyst(s). Finally, cost estimations will also not economically viable [170]. The future of APR is related to these
be affected by external factors, such as the fluctuation of fossil fuel technologies.
prices, the variation in the biofuel policies of a given country, and the Biological processes, widely studied at the pilot scale, have been
carbon tax related to CO2 emissions [160,161]. developed to reach the industrialisation step. The PR and dark fermen­
Table 5 presents the key TRL, CAPEX, and H2 production costs for tation processes are two technologies of interest. Currently, fermenta­
different technologies applied either on biomass, water, or fossil re­ tion processes are more adapted to a small scale. Their high H2
sources (natural gas, coal). Thermochemical processes include tradi­ production costs make them less viable and competitive than thermo­
tional methods from fossil fuels. Currently, steam methane reforming chemical conversion processes. The most significant issue is related to
and coal gasification have the lowest H2 production cost (<1 $/kg H2), the lower yield and rate of biomass conversion [171]. In addition, the
mainly because of the large existing structures (TRL 9) and the low price systems require expensive bioreactors and separation steps to obtain
of the feedstock. However, biomass prices are expected to decrease by pure H2 from gaseous products. This equipment increases the CAPEX of
2050 at the expense of fossil resources. The gasification and pyrolysis of the process, hindering development on an industrial scale [172]. To
biomass are similar to those of technologies performed on fossil feed­ improve the fermentation and enrich the hydrogen fraction,
stock. They are expected to mature to reach a TRL up to 9 in the next two pre-treatments can be performed on recalcitrant biomass. Physical
decades. The critical challenges are related to the type of feedstock, cost pre-treatments, including grinding, chipping, and milling, are the most
of the reactor, and CCS systems combined with the processes. In the case investigated step that are prior to biomass conversion. However, this
of biomass gasification, the reactor represents up to 42.9% of the additional step increases the complexity of the process and can produce
CAPEX. In addition, Bartels et al. demonstrated that large plants with inhibitors of bacterial growth, thereby impacting both the CAPEX and
CCS systems have higher CAPEX. For the SMR process, the integration of the OPEX [173]. Future improvements should focus on the development
CCS systems represents an increase of 25–30% for the H2 production of local small-scale production plants. In addition, the optimisation of
cost. The size of the plant will also significantly affect the H2 production bioreactors and development of new bacterial strains are required.
cost. For example, a small gasification plant will produce H2 for 10 $/kg Regarding electrochemical processes, water electrolysis technologies
compared with a large plant, for which the cost decreases to 1.2.3.5 have a TRL as high as fossil fuel-based processes. However, industrial
$/kg. OPEX accounts for 60% of the CAPEX of a gasification plant and production is currently performed at a smaller scale. Felgenhaur and
60% of the total production cost. Approximately 60% of the OPEX itself Hamacher compared the process efficiencies [174], and determined that
is imputed to the maintenance and insurance, followed by the feedstock for a similar productivity (90 kg/h), the CAPEX of water electrolysis (4
price (25%) and the operational cost of the entire process (10%; energy M€) is lower than gasification (9.9 M€) and steam reforming (11 M€).
and electricity) [162]. The choice of a suitable catalyst will influence However, the investment cost can be substantial for large-scale

Table 5
Comparison of the Technology Readiness Level (TRL), CAPEX and H2 production costs of different processes.
Process TRL Production scale CAPEX H2 production cost ($/kg)12 Ref.

FOSSIL RESOURCES
SMR 9 Large/Available 170.95–240.20 M€1 0.77 [21,80,161]
Coal Gasification 9 Large/Available 257.60 M€2 0.92–2.83 [21,80,164]
WATER
Water Electrolysis 9 Small/Available 4.0 M€9 2.35–4.80 (Nuclear) [21,164,165]
8.9 M€10 5.00–10.00 (RSE13)
Vegetal/algal biomass
Gasification 7 Mid-size/Available 11 M€3 1.21–3.5 [21,61,67]
215.3 M$4
Steam Reforming 8 Small/Available 9.9 M€5 1.83–2.35 [61,166]
Pyrolysis 7 Mid-size/Available 210–287 M$4 1.21–2.57 [21,80,167]
ScWG 4 Pilot plant 57.44 M€6 1.51–3.89 [168]
277.8 M$4
APR 4–5 Pilot plant 12.85 M$7 4.00 [169]
Dark fermentation 5 Pilot plant / 2.57–2.80 [21]
Photo-fermentation 4 Under research 115.6 M$8 2.83–3.89 [21,80]
MEC 2–4 Under research 2.8 M€11 1.7–4.5114 [123]

10
T. Lepage et al. Biomass and Bioenergy 144 (2021) 105920

production due to the price of an electrolyser, which can represent 60% between conventional, hybrid, electric, hydrogen fuelled internal com­
of the CAPEX. According to the CAPEX analysis, electrolysis is currently bustion engines, and hydrogen fuel cells.
more adapted to a small-scale approach (including the case of a Hydrogen-fuelled vehicles consume roughly the same consumption
hydrogen refuelling station) [175]. Among all the processes listed in as hybrid vehicles in the case of HICE and consume more fuel than
Table 5, sustainable water electrolysis presents the highest costs. Elec­ electric vehicles, but always less than that of fossil fuel vehicles. Electric
trical consumption is the primary reason for the high OPEX, which in­ vehicles have the lowest fuel price, followed by hydrogen-fuelled (1.69
creases the H2 production cost, especially when renewable electricity is for HICE and 1.72 for HFC) and hybrid vehicles 1.71. However, con­
used. Yao et al. showed that electrical consumption can represent up to ventional vehicles have the highest fuel price. The fuel dividing range
80% of the operational annual cost (4.8 M€/a) for alkaline electrolysis was approximately 350 km for hydrogen-fuelled vehicles, which is
(with a CAPEX of 4 M€) [61]. The H2 production cost is expected to almost double the distance that electric vehicles can travel. However, it
decrease in the future with the development of current technologies is lower than the fuel dividing range of conventional and hybrid vehi­
from 5.0 €/kg H2 to 2.3 €/kg H2 for alkaline electrolysis [176]. This has cles, which showed the highest dividing range [5,185–187].
motivated the development of cheaper processes from renewable ma­ Concerning the environmental impact, two major criteria are
terials, including MEC, PEMEC, and CEC. Indeed, the current research considered: greenhouse gas (GHG) emissions, particularly CO2 gases
on MEC allows the production of hydrogen at a lower cost (1.7–4.51 correlated to global warming potential, and acidification potential (AP),
$/kg) compared with water electrolysis (2.35–4.8 $/kg). With the essentially SO2 emissions, related to acid rains. Nanaki et al. reported
development of CEC technologies, which have the lowest electrical that (HICE) and (HFC) have the lowest GHG emissions (6.40 kg CO2/
consumption (equivalent to 16.7% of that of water electrolysis), OPEX 100 km of vehicle travel and 7.10 kg CO2/100 km of vehicle travel,
could be drastically reduced as well as the H2 production cost [106]. In respectively) compared with conventional, hybrid, and electric vehicles.
this case, it could also promote the installation of larger plants to Regarding acidification potential, HICE and HFC also had the lowest AP
decrease the hydrogen production cost. The chemical electrolysis con­ (0.02 kg SO2/100 km travel and 0.018 kg SO2/100 km travel, respec­
version of biomass could provide a new method for industrial applica­ tively) compared with other vehicle types that oscillated between 0.04
tions to overcome the high electrical consumption of water electrolysis. kg SO2/100 km travel and 0.06 kg SO2/100 km travel [188].
CEC methods benefit from water electrolysis and are promising for areas Literature data proves that hydrogen-fuelled vehicles including in­
where biomass and waste are abundant. These two techniques comple­ ternal combustion engines and fuel cells are the cleanest because they
ment each other based on the local abundance of biomass, water, or represent the lowest CO2 and SO2 emissions. Fuel cells have the best
green electricity production. Nevertheless, future studies are necessary environmental performance and provide a sufficient dividing range.
for this novel process with a low TRL to predict its impact on the future Hydrogen fuel is promising, eco-friendly, and could be competitive with
of hydrogen. More research is required to develop biomass CEC methods electric vehicles, especially when the production processes are improved
to convert more complex feedstock into hydrogen and to assess its in order to minimise its cost.
economic feasibility.
7. Conclusions
6. Hydrogen assessment as fuel in transportation
Among the initiatives currently taken to partially overcome global
As a future energy, hydrogen is a potential candidate to ensure a warming, a great deal of research has been conducted on hydrogen as an
renewable, sustainable, and alternative energy vector. Various conversion pathways have already
secure fuel [177]. Compared with traditional fuels, H2 has the been investigated to produce H2 from lignocellulosic biomass.
characteristic of being used in both fuel cells and internal combustion Thermochemical processes are the most common technologies.
engines [5]. In fuel cells (HFC), electricity is generated through a com­ Gasification is the predominant process globally and should reach the
bination of hydrogen and oxygen. This electrochemical reaction is industrial scale (TRL 9) in a few decades. Steam gasification is the best
assured by a catalyst that splits H2 into electrons and protons. The compromise as it does not require an expensive oxidising agent (oxygen)
positively charged protons cross the cathode, and the negatively charged and is adapted to a wide variety of wet and dried feedstocks. Typically,
protons are pushed through the system to generate electricity [178]. In increasing the size of the plant will increase the CAPEX but will signif­
internal combustion engines (HICE), hydrogen can be either sent to the icantly reduce the H2 production cost. OPEX is primarily due to main­
manifold as cold hydrogen or be stored in a cryogenic cylinder as liquid tenance, insurance, and the feedstock price. The major drawback
hydrogen and then transformed into a cold gas though the heat remains the tar and char formation and the deactivation of the catalyst.
exchanger or it can be used as an alternative fuel to gasoline to form a Current research is predominantly focused on the use of novel catalysts,
hydrogen-gasoline mixture, which is compressed and ignited by a mainly Ni-based catalysts, biochar, dolomite, and olivine, to improve
combustion spark [5]. the selectivity of reactions and decrease the inhibition issues. Currently,
Hydrogen-fuelled engines have several advantages in terms of min­ the lowest hydrogen production cost from renewable feedstock is met
imum ignition energy, auto-ignition temperature, and flammability. Its with biomass gasification.
minimum ignition energy (the minimum amount of energy required to
ignite the air fuel mixture) is lower than that in gas and gasoline engines,
Table 6
giving it the advantage of rapid ignition [179,180]. In addition, its
Comparison between Hydrogen fuelled vehicles and conventional, hybrid, and
self-ignition requires a higher temperature compared with gas and
electric vehicles [5,185–188].
gasoline, giving it the characteristic of high compression ratios for in­
ternal combustion [180–182]. Its flammability limits in the air (the fuel Vehicles HICE HFC Hybrid Electric Conventional

volumetric concentration limits and the fuel–air mixture flammability) Fuel consumption (MJ/ 136.6 129.50 137.6 67.20 236.8
is wider than that of gas and gasoline, allowing a better fuel economy by 100 km)
Fuel price (USD/100 1.72 1.69 1.71 0.9 2.94
reducing complete combustion [183,184]. In addition, the high diffu­
km)
sivity of hydrogen allows its rapid diffusion in the air, forming a ho­ Fuel dividing range 350 355 930 160 540
mogenous mixture. When there is a leak in the system, hydrogen will be (km)
diffused in the air without harming the system or the environment [10]. GHGa emission 6.40 7.10 13.30 11.40 21.40
In order to evaluate the performance of hydrogen-based vehicles, APb 0.02 0.018 0.04 0.045 0.06

environmental efficiency and fuel assessment were chosen as the a


Kg CO2 eq. per 100 km of travel.
criteria. Table 6 presents a comparison of these selected parameters b
kg SO2 eq. per 100 km of travel.

11
T. Lepage et al. Biomass and Bioenergy 144 (2021) 105920

In addition, studies have been conducted to use microorganisms that [4] M. Raud, T. Kikas, O. Sippula, N.J. Shurpali, Potentials and challenges in
lignocellulosic biofuel production technology, Renew. Sustain. Energy Rev. 111
allow the biological conversion of biomass, mostly through dark
(2019) 44–56, https://doi.org/10.1016/j.rser.2019.05.020.
fermentation processes. The key advantages reside under mild condi­ [5] C. Acar, I. Dincer, The potential role of hydrogen as a sustainable transportation
tions (T◦ and pressure) used. The main limitations of these processes are fuel to combat global warming, Int. J. Hydrogen Energy (in press) (n.d.). https://
the slow conversion rate and the low volumes produced, hindering the doi.org/10.1016/j.ijhydene.2018.10.149.
[6] P. Crane, D.S. Scott, Efficiency and CO2 emission analysis of pathways by which
development (TRL 5). These systems require expensive bioreactors and methane can provide transportation services, Int. J. Hydrogen Energy 17 (1992)
separation steps to increase CAPEX. In addition, pre-treatments are 543–550, https://doi.org/10.1016/0360-3199(92)90154-O.
necessary for recalcitrant biomass, inducing the production of inhibitors [7] P.E. Dodds, I. Staffell, A.D. Hawkes, F. Li, P. Grünewald, W. McDowall, P. Ekins,
Hydrogen and fuel cell technologies for heating: a review, Int. J. Hydrogen
and negatively impacting CAPEX and OPEX. Future improvements Energy 40 (2015) 2065–2083, https://doi.org/10.1016/j.ijhydene.2014.11.059.
should be focused on the development of local small-scale production [8] R.M. Navarro, M.C. Sanchez-Sanchez, M.C. Alvarez-Galvan, J.L.G. Fierro, A.-
plants as well as the development of better-performing bioreactors and Z. Saeed, H2 production from renewables, in: Encycl. Inorg. Chem., John Wiley &
Sons Ltd., 2011, pp. 1–18, https://doi.org/10.1002/9781119951438.eibc0450.
new bacterial strains. In addition, they are more suitable for liquid [9] J.A. Turner, Sustainable hydrogen production, Science 305 (2004) 972–974,
streams containing sugars, such as molasses and wastewater (sewage https://doi.org/10.1126/science.1103197, 80.
sludge) that can be processed by microorganisms. Interest has recently [10] S. Sharma, S.K. Ghoshal, Hydrogen the future transportation fuel: from
production to applications, Renew. Sustain. Energy Rev. 43 (2015) 1151–1158,
increased concerning these feedstocks because they are easily processed https://doi.org/10.1016/j.rser.2014.11.093.
by microorganisms, while they are challenging in major processes such [11] International Energy Agency (IEA), The Future of Hydrogen, Japan, https://doi.
as gasification. org/10.1016/S1464-2859(12)70027-5, 2019.
[12] R.C. Saxena, D. Seal, S. Kumar, H.B. Goyal, Thermo-chemical routes for hydrogen
Among the latest conversion techniques, electrochemical methods,
rich gas from biomass: a review, Renew. Sustain. Energy Rev. 12 (2008)
including biomass electrolysis, deliver promising results in terms of 1909–1927, https://doi.org/10.1016/j.rser.2007.03.005.
yield at the laboratory scale (TRL 2–4) while consuming less electricity [13] E.S. Hanley, J.P. Deane, B.P.Ó. Gallachóir, The role of hydrogen in low carbon
than conventional water electrolysis. Notably, recently developed MEC energy futures–A review of existing perspectives, Renew. Sustain. Energy Rev. 82
(2018) 3027–3045, https://doi.org/10.1016/j.rser.2017.10.034.
allows the production of hydrogen at a similar cost compared with water [14] F. Suleman, I. Dincer, M. Agelin-Chaab, Environmental impact assessment and
electrolysis. With the development of CEC technologies, which have the comparison of some hydrogen production options, Int. J. Hydrogen Energy 40
lowest electrical consumption, OPEX could be drastically reduced as (2015) 6976–6987, https://doi.org/10.1016/j.ijhydene.2015.03.123.
[15] Alexander E. Farrell, R.J. Plevin, B.T. Turner, A.D. Jones, M. O’Hare, D.
well as the H2 production cost. Combined with cheap and clean wind or M. Kammen, Ethanol can contribute to energy and environmental goals, Science
solar energy, H2 could be produced in areas producing large amounts of 80 (311) (2006) 506–509, https://doi.org/10.1126/science.1121416.
biomass. [16] G. Li, P. Cui, Y. Wang, Z. Liu, Z. Zhu, S. Yang, Life cycle energy consumption and
GHG emissions of biomass-to- hydrogen process in comparison with coal-to-
Numerous applications are currently dependent on hydrogen. The hydrogen process, Energy. (n.d.). https://doi.org/10.1016/j.energy.20
current increasing demand is encouraging the boom of a new economy 19.116588.
based on hydrogen use. The conversion of biomass coupled with CSS [17] I. Dincer, Green methods for hydrogen production, Int. J. Hydrogen Energy 37
(2012) 1954–1971, https://doi.org/10.1016/j.ijhydene.2011.03.173.
systems makes it possible to obtain negative CO2 emissions compared [18] Y. Kalinci, A. Hepbasli, I. Dincer, Biomass-based hydrogen production: a review
with the reforming of fossil resources. Biomass conversion still needs and analysis, Int. J. Hydrogen Energy 34 (2009) 8799–8817, https://doi.org/
improvements and adaptations to be implemented at a larger scale, 10.1016/j.ijhydene.2009.08.078.
[19] H. Balat, E. Kirtay, Hydrogen from biomass - present scenario and future
allowing the production of larger volumes to be competitive, especially
prospects, Int. J. Hydrogen Energy 35 (2010) 7416–7426, https://doi.org/
in the biofuel sector. 10.1016/j.ijhydene.2010.04.137.
[20] S.E. Hosseini, M.A. Wahid, M.M. Jamil, A.A.M. Azli, M.F. Misbah, A review on
Supplementary information biomass-based hydrogen production for renewable energy supply, Int. J. Energy
Res. 39 (2015) 1597–1615, https://doi.org/10.1002/er.3381.
[21] P. Parthasarathy, K.S. Narayanan, Hydrogen production from steam gasification
This paper is associated with the S1 document showing, in table of biomass: influence of process parameters on hydrogen yield - a review, Renew.
form, the H2 yields from various biomasses, estimated in g per kg of dry Energy 66 (2014) 570–579, https://doi.org/10.1016/j.renene.2013.12.025.
[22] W. Liu, J. Wang, D. Bhattacharyya, Y. Jiang, D. DeVallance, Economic and
biomass. environmental analyses of coal and biomass to liquid fuels, Energy 141 (2017)
76–86, https://doi.org/10.1016/j.energy.2017.09.047.
Acknowledgments [23] G. Fiorese, M. Catenacci, V. Bosetti, E. Verdolini, The power of biomass: experts
disclose the potential for success of bioenergy technologies, Energy Pol. 65 (2014)
94–114, https://doi.org/10.1016/j.enpol.2013.10.015.
This research was supported by the European Union and the Walloon [24] M. Kammoun, H. Ayeb, T. Bettaieb, A. Richel, Chemical characterisation and
Region with the European Funds for Regional Development 2014–2020 technical assessment of agri-food residues, marine matrices, and wild grasses in
the South Mediterranean area: a considerable inflow for biorefineries, Waste
in the framework of the VERDIR Tropical Plant Factory program (project Manag. 118 (2020) 247–257, https://doi.org/10.1016/j.wasman.2020.08.032.
BioResidu). [25] N. Dahmen, I. Lewandowski, S. Zibek, A. Weidtmann, Integrated lignocellulosic
value chains in a growing bioeconomy : status quo and perspectives, Bioenergy
(2019) 107–117, https://doi.org/10.1111/gcbb.12586.
Appendix A. Supplementary data
[26] V. Stenberg, M. Rydén, T. Mattisson, A. Lyngfelt, Exploring novel hydrogen
production processes by integration of steam methane reforming with chemical-
Supplementary data to this article can be found online at https://doi. looping combustion (CLC-SMR) and oxygen carrier aided combustion (OCAC-
org/10.1016/j.biombioe.2020.105920. SMR), Int. J. Greenh. Gas Control. 74 (2018) 28–39, https://doi.org/10.1016/j.
ijggc.2018.01.008.
[27] G.W. Huber, S. Iborra, A. Corma, Synthesis of transportation fuels from biomass:
References chemistry, catalysts, and engineering, Chem. Rev. 106 (2006) 4044–4098,
https://doi.org/10.1021/cr068360d.
[1] P. Capros, G. Zazias, S. Evangelopoulou, M. Kannavou, T. Fotiou, P. Siskos, A. De [28] C. Xu, S. Chen, A. Soomro, Z. Sun, W. Xiang, Hydrogen rich syngas production
Vita, K. Sakellaris, Energy-system modelling of the EU strategy towards climate- from biomass gasification using synthesized Fe/CaO active catalysts, J. Energy
neutrality, Energy Pol. 134 (2019) 110960, https://doi.org/10.1016/j. Inst. 91 (2018) 805–816, https://doi.org/10.1016/j.joei.2017.10.014.
enpol.2019.110960. [29] C. Srinivasakannan, N. Balasubramanian, Variations in the design of dual
[2] S.E. Hosseini, M.A. Wahid, Hydrogen production from renewable and sustainable fluidized bed gasifiers and the quality of syngas from biomass, Energy Sources,
energy resources: promising green energy carrier for clean development, Renew. Part A Recover, Util. Environ. Eff. 33 (2011) 349–359, https://doi.org/10.1080/
Sustain. Energy Rev. 57 (2016) 850–866, https://doi.org/10.1016/j. 15567030902967835.
rser.2015.12.112. [30] A. Adamovics, R. Platace, I. Gulbe, S. Ivanovs, The content of carbon and
[3] X.-Y. Li, B.-J. Tang, Incorporating the transport sector into carbon emission hydrogen in grass biomass and its influence on heating value, Eng. Rural Dev. 17
trading scheme: an overview and outlook, Nat. Hazards 88 (2017) 683–698, (2018) 1277–1281, https://doi.org/10.22616/ERDev2018.17.N014.
https://doi.org/10.1007/s11069-017-2886-3. [31] T.G. Madenoǧlu, M. Saǧlam, M. Yüksel, L. Ballice, Hydrothermal gasification of
biomass model compounds (cellulose and lignin alkali) and model mixtures,

12
T. Lepage et al. Biomass and Bioenergy 144 (2021) 105920

J. Supercrit. Fluids 115 (2016) 79–85, https://doi.org/10.1016/j. [57] A.V. Bridgwater, Renewable fuels and chemicals by thermal processing of
supflu.2016.04.017. biomass, Chem. Eng. J. 91 (2003) 87–102, https://doi.org/10.1016/S1385-8947
[32] J. Gil, M.P. Aznar, M.A. Caballero, E. Francés, J. Corrella, Biomass gasification in (02)00142-0.
fluidized bed at pilot scale with steam-oxygen mixtures. Product distribution for [58] L. Barelli, G. Bidini, F. Gallorini, S. Servili, Hydrogen production through
very different operating condictions, Energy Fuels 11 (1997) 1109–1118, https:// sorption-enhanced steam methane reforming and membrane technology: a
doi.org/10.1021/ef9602335. review, Energy 33 (2008) 554–570, https://doi.org/10.1016/j.
[33] Z. Wang, T. He, J. Qin, J. Wu, J. Li, Z. Zi, G. Liu, J. Wu, L. Sun, Gasification of energy.2007.10.018.
biomass with oxygen-enriched air in a pilot scale two-stage gasifier, Fuel 150 [59] D.P. Harrison, Sorption-enhanced hydrogen Production : a review, Ind. Eng.
(2015) 386–393, https://doi.org/10.1016/j.fuel.2015.02.056. Chem. Res. 47 (2008) 6486–6501, https://doi.org/10.1021/ie800298z.
[34] S. Rapagná, H. Provendier, C. Petit, A. Kiennemann, P.U. Foscolo, Development of [60] S. Kurella, P.K. Bhukya, B.C. Meikap, Removal of H2S pollutant from gasifier
catalysts suitable for hydrogen or syn-gas production from biomass gasification, syngas by a multistage dual-flow sieve plate column wet scrubber, J. Environ. Sci.
Biomass Bioenergy 22 (2002) 377–388, https://doi.org/10.1016/S0961-9534 Heal. Part A. 52 (2017) 515–523, https://doi.org/10.1080/
(02)00011-9. 10934529.2017.1281690.
[35] P. Basu, Gasification theory and modeling of gasifiers, in: Biomass Gasif. Des. [61] J. Yao, M. Kraussler, F. Benedikt, H. Hofbauer, Techno-economic assessment of
Handb., First Edit, © 2010 Elsevier Inc., 2010, pp. 117–165, https://doi.org/ hydrogen production based on dual fluidized bed biomass steam gasification,
10.1016/b978-0-12-374988-8.00005-2. biogas steam reforming, and alkaline water electrolysis processes, Energy
[36] H. Yang, H. Chen, Biomass gasification for synthetic liquid fuel production, in: Convers. Manag. 145 (2017) 278–292, https://doi.org/10.1016/j.
Gasif. Synth. Fuel Prod. Fundam. Process. Appl., ©, Woodhead Publishing enconman.2017.04.084.
Limited. All rights reserved., 2015, pp. 241–275, https://doi.org/10.1016/B978- [62] C. Pirez, W. Fang, M. Capron, S. Paul, H. Jobic, F. Dumeignil, L. Jalowiecki-
0-85709-802-3.00011-4, 2015. Duhamel, Steam reforming, partial oxidation and oxidative steam reforming for
[37] H. De Lasa, E. Salaices, J. Mazumder, R. Lucky, Catalytic steam gasification of hydrogen production from ethanol over cerium nickel based oxyhydride catalyst,
biomass: catalysts, thermodynamics and kinetics, Chem. Rev. 111 (2011) Appl. Catal. Gen. 518 (2016) 78–86, https://doi.org/10.1016/j.
5404–5433, https://doi.org/10.1021/cr200024w. apcata.2015.10.035.
[38] Pekka Simell, E. Kurkela, P. Ståhlberg, J. Hepola, Catalytic hot gas cleaning of [63] D. Rennard, R. French, S. Czernik, T. Josephson, L. Schmidt, Production of
gasification gas, Catal. Today 27 (1996) 55–62, https://doi.org/10.1016/S0140- synthesis gas by partial oxidation and steam reforming of biomass pyrolysis oils,
6701(97)83950-1. Int. J. Hydrogen Energy 35 (2010) 4048–4059, https://doi.org/10.1016/j.
[39] A. Erkiaga, G. Lopez, M. Amutio, J. Bilbao, M. Olazar, Steam gasification of ijhydene.2010.01.143.
biomass in a conical spouted bed reactor with olivine and γ-alumina as primary [64] J.R. Salge, G.A. Deluga, L.D. Schmidt, Catalytic partial oxidation of ethanol over
catalysts, Fuel Process, Technol. 116 (2013) 292–299, https://doi.org/10.1016/j. noble metal catalysts, J. Catal. 235 (2005) 69–78, https://doi.org/10.1016/j.
fuproc.2013.07.008. jcat.2005.07.021.
[40] I. Ahmed, A.K. Gupta, Evolution of syngas from cardboard gasification, Appl. [65] M. Tóth, E. Varga, A. Oszkó, K. Baán, J. Kiss, A. Erdohelyi, Partial oxidation of
Energy 86 (2009) 1732–1740, https://doi.org/10.1016/j.apenergy.2008.11.018. ethanol on supported Rh catalysts: effect of the oxide support, J. Mol. Catal.
[41] H. Yang, R. Yan, H. Chen, D.H. Lee, D.T. Liang, C. Zheng, Pyrolysis of palm oil Chem. 411 (2016) 377–387, https://doi.org/10.1016/j.molcata.2015.11.010.
wastes for enhanced production of hydrogen rich gases, Fuel Process. Technol. 87 [66] L. He, J. Yang, D. Chen, Hydrogen from biomass: advances in thermochemical
(2006) 935–942, https://doi.org/10.1016/j.fuproc.2006.07.001. processes, in: Renew. Hydrog. Technol. Prod., Purif., Storage, Appl. Saf., Elsevier
[42] H.B. Goyal, D. Seal, R.C. Saxena, Bio-fuels from thermochemical conversion of B.V., 2013, pp. 111–133, https://doi.org/10.1016/B978-0-444-56352-1.00006-
renewable resources: a review, Renew. Sustain. Energy Rev. 12 (2008) 504–517, 4.
https://doi.org/10.1016/j.rser.2006.07.014. [67] M. Kumar, A.O. Oyedun, A. Kumar, A comparative analysis of hydrogen
[43] C. Di Blasi, Modeling chemical and physical processes of wood and biomass production from the thermochemical conversion of algal biomass, Int. J.
pyrolysis, Prog. Energy Combust. Sci. 34 (2008) 47–90, https://doi.org/10.1016/ Hydrogen Energy 44 (2019) 10384–10397, https://doi.org/10.1016/j.
j.pecs.2006.12.001. ijhydene.2019.02.220.
[44] L. Garcia, R. French, S. Czernik, E. Chornet, Catalytic steam reforming of bio-oils [68] B.R. Pinkard, D.J. Gorman, K. Tiwari, E.G. Rasmussen, J.C. Kramlich, P.
for the production of hydrogen: effects of catalyst composition, Appl. Catal. Gen. G. Reinhall, I.V. Novosselov, Supercritical water gasification: practical design
201 (2000) 225–239, https://doi.org/10.1016/S0926-860X(00)00440-3. strategies and operational challenges for lab-scale, continuous flow reactors,
[45] C. Font Palma, Modelling of tar formation and evolution for biomass gasification: Heliyon 5 (2019), e01269, https://doi.org/10.1016/j.heliyon.2019.e01269.
a review, Appl. Energy 111 (2013) 129–141, https://doi.org/10.1016/j. [69] L.J. Guo, Y.J. Lu, X.M. Zhang, C.M. Ji, Y. Guan, A.X. Pei, Hydrogen production by
apenergy.2013.04.082. biomass gasification in supercritical water: a systematic experimental and
[46] S. Koppatz, C. Pfeifer, H. Hofbauer, Comparison of the performance behaviour of analytical study, Catal. Today 129 (2007) 275–286, https://doi.org/10.1016/j.
silica sand and olivine in a dual fluidised bed reactor system for steam gasification cattod.2007.05.027.
of biomass at pilot plant scale, Chem. Eng. J. 175 (2011) 468–483, https://doi. [70] M. Osada, T. Sato, M. Watanabe, M. Shirai, K. Arai, Catalytic gasification of wood
org/10.1016/j.cej.2011.09.071. biomass in subcritical and supercritical water, Combust. Sci. Technol. 178 (2006)
[47] A. Demirbaş, Gaseous products from biomass by pyrolysis and gasification: effects 537–552, https://doi.org/10.1080/00102200500290807.
of catalyst on hydrogen yield, Energy Convers. Manag. 43 (2002) 897–909, [71] Y. Matsumura, T. Minowa, B. Potic, S.R.A. Kersten, W. Prins, W.P.M. Van Swaaij,
https://doi.org/10.1016/S0196-8904(01)00080-2. B. Van De Beld, D.C. Elliott, G.G. Neuenschwander, A. Kruse, M.J. Antal Jr.,
[48] E. Kirtay, Recent advances in production of hydrogen from biomass, Energy Biomass gasification in near- and super-critical water: status and prospects,
Convers. Manag. 52 (2011) 1778–1789, https://doi.org/10.1016/j. Biomass Bioenergy 29 (2005) 269–292, https://doi.org/10.1016/j.
enconman.2010.11.010. biombioe.2005.04.006.
[49] M. Van de Velden, J. Baeyens, A. Brems, B. Janssens, R. Dewil, Fundamentals, [72] A.B.H. Trabelsi, A. Ghrib, K. Zaafouri, A. Friaa, A. Ouerghi, S. Naoui,
kinetics and endothermicity of the biomass pyrolysis reaction, Renew. Energy 35 H. Belayouni, Hydrogen-rich syngas production from gasification and pyrolysis of
(2010) 232–242, https://doi.org/10.1016/j.renene.2009.04.019. solar dried sewage sludge: experimental and modeling investigations, BioMed
[50] F. Safari, M. Salimi, A. Tavasoli, A. Ataei, Non-catalytic conversion of wheat Res. Int. (2017) 1–14, https://doi.org/10.1155/2017/7831470.
straw, walnut shell and almond shell into hydrogen rich gas in supercritical water [73] A.V. Bridgwater, A.J. Toft, J.G. Brammer, in: A Techno-Economic Comparison of
media, Chin. J. Chem. Eng. 24 (2016) 1097–1103, https://doi.org/10.1016/j. Power Production by Biomass Fast Pyrolysis with Gasification and Combustion,
cjche.2016.03.002. 2002, https://doi.org/10.1016/S1364-0321(01)00010-7.
[51] P.M. Lv, Z.H. Xiong, J. Chang, C.Z. Wu, Y. Chen, J.X. Zhu, An experimental study [74] D. Mohan, C.U. Pittman, P.H. Steele, Pyrolysis of wood/biomass for bio-oil: a
on biomass air-steam gasification in a fluidized bed, Bioresour. Technol. 95 critical review, Energy Fuels 20 (2006) 848–889, https://doi.org/10.1021/
(2004) 95–101, https://doi.org/10.1016/j.biortech.2004.02.003. ef0502397.
[52] A.C.C. Chang, H.-F. Chang, F.-J. Lin, K.-H. Lin, C.-H. Chen, Biomass gasification [75] A. Demirbas, Comparison of thermochemical conversion processes of biomass to
for hydrogen production, Int. J. Hydrogen Energy 36 (2011) 14252–14260, hydrogen-rich gas mixtures, Energy Sources, Part A Recover, Util. Environ. Eff. 38
https://doi.org/10.1016/j.ijhydene.2011.05.105. (2016) 2971–2976, https://doi.org/10.1080/15567036.2015.1122686.
[53] A. Kumar, K. Eskridge, D.D. Jones, M.A. Hanna, Steam-air fluidized bed [76] M. Balat, Hydrogen-rich gas production from biomass via pyrolysis and
gasification of distillers grains: effects of steam to biomass ratio, equivalence ratio gasification processes and effects of catalyst on hydrogen yield, Energy Sources,
and gasification temperature, Bioresour. Technol. 100 (2009) 2062–2068, Part A. 30 (2008) 552–564, https://doi.org/10.1080/15567030600817191.
https://doi.org/10.1016/j.biortech.2008.10.011. [77] M. Ni, D.Y.C. Leung, M.K.H. Leung, K. Sumathy, An overview of hydrogen
[54] A. Arregi, M. Amutio, G. Lopez, J. Bilbao, M. Olazar, Evaluation of production from biomass, Fuel Process, Technol. 87 (2006) 461–472, https://doi.
thermochemical routes for hydrogen production from biomass: a review, Energy org/10.1016/j.fuproc.2005.11.003.
Convers. Manag. 165 (2018) 696–719, https://doi.org/10.1016/j. [78] P. Lan, Q. Xu, M. Zhou, L. Lan, S. Zhang, Y. Yan, Catalytic steam reforming of fast
enconman.2018.03.089. pyrolysis bio-oil in fixed bed and fluidized bed reactors, Chem. Eng. Technol. 33
[55] L. Devi, K.J. Ptasinski, F.J.J.G. Janssen, S.V.B. van Paasen, P.C.A. Bergman, J.H. (2010) 2021–2028, https://doi.org/10.1002/ceat.201000169.
A. Kiel, Catalytic decomposition of biomass tars: use of dolomite and untreated [79] S. Czernik, R. Evans, R. French, Hydrogen from biomass-production by steam
olivine, Renew. Energy 30 (2005) 565–587, https://doi.org/10.1016/j. reforming of biomass pyrolysis oil, Catal. Today 129 (2007) 265–268, https://doi.
renene.2004.07.014. org/10.1016/j.cattod.2006.08.071.
[56] Y. Lu, S. Li, L. Guo, X. Zhang, Hydrogen production by biomass gasification in [80] V. Martinez-Merino, M.J. Gil, A. Cornejo, Biomass sources for hydrogen
supercritical water over Ni/γAl2O3 and Ni/CeO2-γAl2O3 catalysts, Int. J. production, in: Renew. Hydrog. Technol. Prod., Purif., Storage, Appl. Saf.,
Hydrogen Energy 35 (2010) 7161–7168, https://doi.org/10.1016/j. Elsevier B.V., 2013, pp. 87–110, https://doi.org/10.1016/B978-0-444-56352-
ijhydene.2009.12.047. 1.00005-2.

13
T. Lepage et al. Biomass and Bioenergy 144 (2021) 105920

[81] G.W. Huber, J.A. Dumesic, An overview of aqueous-phase catalytic processes for systems (BESs): technology for sustainable electricity, waste remediation,
production of hydrogen and alkanes in a biorefinery, Catal. Today 111 (2006) resource recovery, chemical production and beyond, Renew. Energy 98 (2016)
119–132, https://doi.org/10.1016/j.cattod.2005.10.010. 153–170, https://doi.org/10.1016/j.renene.2016.03.002.
[82] R.D. Cortright, R.R. Davda, J.A. Dumesic, Hydrogen from catalytic reforming of [106] W. Liu, Y. Cui, X. Du, Z. Zhang, Z. Chao, Y. Deng, High efficiency hydrogen
biomass-derived hydrocarbons in liquid water, Nature 418 (2002) 964–967, evolution from native biomass electrolysis, Energy Environ. Sci. 9 (2016)
https://doi.org/10.1038/nature01009. 467–472, https://doi.org/10.1039/C5EE03019F.
[83] A. Tanksale, J.N. Beltramini, G.M. Lu, A review of catalytic hydrogen production [107] Y.X. Chen, A. Lavacchi, H.A. Miller, M. Bevilacqua, J. Filippi, M. Innocenti,
processes from biomass, Renew. Sustain. Energy Rev. 14 (2010) 166–182, A. Marchionni, W. Oberhauser, L. Wang, F. Vizza, Nanotechnology makes
https://doi.org/10.1016/j.rser.2009.08.010. biomass electrolysis more energy efficient than water electrolysis, Nat. Commun.
[84] R.R. Davda, J.W. Shabaker, G.W. Huber, R.D. Cortright, J.A. Dumesic, A review of 5 (2014) 1–6, https://doi.org/10.1038/ncomms5036.
catalytic issues and process conditions for renewable hydrogen and alkanes by [108] M. Shahbaz, S. Yusup, A. Inayat, D.O. Patrick, M. Ammar, The influence of
aqueous-phase reforming of oxygenated hydrocarbons over supported metal catalysts in biomass steam gasification and catalytic potential of coal bottom ash
catalysts, Appl. Catal. B Environ. 56 (2005) 171–186, https://doi.org/10.1016/j. in biomass steam gasification: a review, Renew. Sustain. Energy Rev. 73 (2017)
apcatb.2004.04.027. 468–476, https://doi.org/10.1016/j.rser.2017.01.153.
[85] D.B. Levin, L. Pitt, M. Love, Biohydrogen production: prospects and limitations to [109] I. Dincer, C. Acar, Review and evaluation of hydrogen production methods for
practical application, Int. J. Hydrogen Energy 29 (2004) 173–185, https://doi. better sustainability, Int. J. Hydrogen Energy 40 (2015) 11094–11111, https://
org/10.1016/S0360-3199(03)00094-6. doi.org/10.1016/j.ijhydene.2014.12.035.
[86] H.H. Larsen, R.K. Feidenhans’l, L. Sønderberg Petersen, Risø Energy Report 3. [110] Y. Wang, S. Wang, G. Zhao, Y. Guo, Y. Guo, Hydrogen production by partial
Hydrogen and its Competitors, Risø National Laboratory, Roskilde, 2004. oxidation gasification of a phenol, naphthalene, and acetic acid mixture in
[87] A.M. Henstra, J. Sipma, A. Rinzema, A.J. Stams, Microbiology of synthesis gas supercritical water, Int. J. Hydrogen Energy 41 (2016) 2238–2246, https://doi.
fermentation for biofuel production, Curr. Opin. Biotechnol. 18 (2007) 200–206, org/10.1016/j.ijhydene.2015.12.115.
https://doi.org/10.1016/j.copbio.2007.03.008. [111] Q. Guan, C. Wei, X. Chai, P. Ning, S. Tian, J. Gu, Q. Chen, R. Miao, Energetic
[88] Y. Liu, R. Lin, Y. Man, J. Ren, Recent developments of hydrogen production from analysis of gasification of biomass by partial oxidation in supercritical water,
sewage sludge by biological and thermochemical process, Int. J. Hydrogen Energy Chin. J. Chem. Eng. 23 (2015) 205–212, https://doi.org/10.1016/j.
44 (2019), https://doi.org/10.1016/j.ijhydene.2019.06.044, 19676–19697. cjche.2014.10.001.
[89] S.S. Hosseini, M. Aghbashlo, M. Tabatabaei, H. Younesi, G. Najafpour, Exergy [112] A.M. Abdalla, S. Hossain, O.B. Nisfindy, A.T. Azad, M. Dawood, A.K. Azad,
analysis of biohydrogen production from various carbon sources via anaerobic Hydrogen production, storage, transportation and key challenges with
photosynthetic bacteria (Rhodospirillum rubrum), Energy 93 (2015) 730–739, applications: a review, Energy Convers. Manag. 165 (2018) 602–627, https://doi.
https://doi.org/10.1016/j.energy.2015.09.060. org/10.1016/j.enconman.2018.03.088.
[90] M. Alfano, C. Cavazza, The biologically mediated water–gas shift reaction: [113] G. Wen, Y. Xu, Z. Xu, Z. Tian, Direct conversion of cellulose into hydrogen by
structure, function and biosynthesis of monofunctional [NiFe]-carbon monoxide aqueous-phase reforming process, Catal. Commun. 11 (2010) 522–526, https://
dehydrogenases, Sustain. Energy Fuels. 2 (2018) 1653–1670, https://doi.org/ doi.org/10.1016/j.catcom.2009.12.008.
10.1039/C8SE00085A. [114] B. Zhao, X. Zhang, L. Sun, G. Meng, L. Chen, Y. Xiaolu, Hydrogen production from
[91] P.C. Hallenbeck, J.R. Benemann, Biological hydrogen production; fundamentals biomass combining pyrolysis and the secondary decomposition, Int. J. Hydrogen
and limiting processes, Int. J. Hydrogen Energy 27 (2002) 1185–1193, https:// Energy 35 (2010) 2606–2611, https://doi.org/10.1016/j.ijhydene.2009.04.011.
doi.org/10.1016/S0360-3199(02)00131-3. [115] J. Wang, Y. Yin, Fermentative hydrogen production using pretreated microalgal
[92] F.R. Hawkes, R. Dinsdale, D.L. Hawkes, I. Hussy, Sustainable fermentative biomass as feedstock, Microb. Cell Factories 17 (2018) 1–16, https://doi.org/
hydrogen production: challenges for process optimisation, Int. J. Hydrogen 10.1186/s12934-018-0871-5.
Energy 27 (2002) 1339–1347, https://doi.org/10.1016/S0360-3199(02)00090- [116] P. Kaparaju, M. Serrano, A.B. Thomsen, P. Kongjan, I. Angelidaki, Bioethanol,
3. biohydrogen and biogas production from wheat straw in a biorefinery concept,
[93] G. Kumar, S. Shobana, D. Nagarajan, D.-J. Lee, K.-S. Lee, C.-Y. Lin, C.-Y. Chen, J.- Bioresour. Technol. 100 (2009) 2562–2568, https://doi.org/10.1016/j.
S. Chang, Biomass based hydrogen production by dark fermentation — recent biortech.2008.11.011.
trends and opportunities for greener processes, Curr. Opin. Biotechnol. 50 (2018) [117] A. Adessi, M. Venturi, F. Candeliere, V. Galli, L. Granchi, R. De Philippis, Bread
136–145, https://doi.org/10.1016/j.copbio.2017.12.024. wastes to energy: sequential lactic and photo-fermentation for hydrogen
[94] S. Eker, M. Sarp, Hydrogen gas production from waste paper by dark production, Int. J. Hydrogen Energy 43 (2018) 9569–9576, https://doi.org/
fermentation: effects of initial substrate and biomass concentrations, Int. J. 10.1016/j.ijhydene.2018.04.053.
Hydrogen Energy 42 (2017) 2562–2568, https://doi.org/10.1016/j. [118] A. Kadier, M.S. Kalil, P. Abdeshahian, K. Chandrasekhar, A. Mohamed, N.
ijhydene.2016.04.020. F. Azman, W. Logroño, Y. Simayi, A.A. Hamid, Recent advances and emerging
[95] A. Ghimire, L. Frunzo, F. Pirozzi, E. Trably, R. Escudie, P.N.L. Lens, G. Esposito, challenges in microbial electrolysis cells (MECs) for microbial production of
A review on dark fermentative biohydrogen production from organic biomass: hydrogen and value-added chemicals, Renew. Sustain. Energy Rev. 61 (2016)
process parameters and use of by-products, Appl. Energy 144 (2015) 73–95, 501–525, https://doi.org/10.1016/j.rser.2016.04.017.
https://doi.org/10.1016/j.apenergy.2015.01.045. [119] J. Sachs, S. Hidayat, S. Giarola, A. Hawkes, in: The Role of CCS and Biomass-
[96] P.C. Hallenbeck, Y. Liu, Recent advances in hydrogen production by Based Processes in the Refinery Sector for Different Carbon Scenarios, Elsevier
photosynthetic bacteria, Int. J. Hydrogen Energy 41 (2016) 4446–4454, https:// Masson SAS, 2018, https://doi.org/10.1016/B978-0-444-64235-6.50239-4.
doi.org/10.1016/j.ijhydene.2015.11.090. [120] V.S. Sikarwar, M. Zhao, P.S. Fennell, N. Shah, E.J. Anthony, Progress in biofuel
[97] E. Sagir, E. Ozgur, U. Gunduz, I. Eroglu, M. Yucel, Single-stage photofermentative production from gasification, Prog. Energy Combust. Sci. 61 (2017) 189–248,
biohydrogen production from sugar beet molasses by different purple non-sulfur https://doi.org/10.1016/j.pecs.2017.04.001.
bacteria, Bioproc. Biosyst. Eng. 40 (2017) 1589–1601, https://doi.org/10.1007/ [121] V. Singh Yadav, R. Vinoth, D. Yadav, Bio-hydrogen production from waste
s00449-017-1815-x. materials: a review, MATEC Web Conf 192 (2018) 1–5, https://doi.org/10.1051/
[98] R. Zagrodnik, M. Laniecki, The role of pH control on biohydrogen production by matecconf/201819202020.
single stage hybrid dark- and photo-fermentation, Bioresour. Technol. 194 (2015) [122] M.M. Rashid, M.K. Al Mesfer, H. Naseem, M. Danish, Hydrogen production by
187–195, https://doi.org/10.1016/j.biortech.2015.07.028. water electrolysis: a review of alkaline water electrolysis, PEM water electrolysis
[99] S. Ghosh, U.K. Dairkee, R. Chowdhury, P. Bhattacharya, Hydrogen from food and high temperature water electrolysis, Int. J. Eng. Adv. Technol. 4 (2015)
processing wastes via photofermentation using Purple Non-sulfur Bacteria (PNSB) 80–93.
– a review, Energy Convers. Manag. 141 (2017) 299–314, https://doi.org/ [123] A. Kadier, M.S. Kalil, A. Mohamed, H.A. Hasan, P. Abdeshahian, T. Fooladi, A.
10.1016/j.enconman.2016.09.001. A. Hamid, Microbial electrolysis cells (MECs) as innovative technology for
[100] H.H. Al-Mohammedawi, H. Znad, E. Eroglu, Improvement of photofermentative sustainable hydrogen production: fundamentals and perspective applications,
biohydrogen production using pre-treated brewery wastewater with banana peels Hydrog. Prod. Technol. (2017) 407–457, https://doi.org/10.1002/
waste, Int. J. Hydrogen Energy 44 (2019) 2560–2568, https://doi.org/10.1016/j. 9781119283676.ch11.
ijhydene.2018.11.223. [124] D.C. Aiken, T.P. Curtis, E.S. Heidrich, Avenues to the financial viability of
[101] M.Y. Azwar, M.A. Hussain, A.K. Abdul-Wahab, Development of biohydrogen microbial electrolysis cells [MEC] for domestic wastewater treatment and
production by photobiological, fermentation and electrochemical processes: a hydrogen production, Int. J. Hydrogen Energy 44 (2019) 2426–2434, https://doi.
review, Renew. Sustain. Energy Rev. 31 (2014) 158–173, https://doi.org/ org/10.1016/j.ijhydene.2018.12.029.
10.1016/j.rser.2013.11.022. [125] M.A. Hamad, A.M. Radwan, D.A. Heggo, T. Moustafa, Hydrogen rich gas
[102] R. Kothari, D. Buddhi, R.L. Sawhney, Comparison of environmental and economic production from catalytic gasification of biomass, Renew. Energy 85 (2016)
aspects of various hydrogen production methods, Renew. Sustain. Energy Rev. 12 1290–1300, https://doi.org/10.1016/j.renene.2015.07.082.
(2008) 553–563, https://doi.org/10.1016/j.rser.2006.07.012. [126] B. Zhang, L. Zhang, Z. Yang, Y. Yan, G. Pu, M. Guo, Hydrogen-rich gas production
[103] A. Kadier, Y. Simayi, P. Abdeshahian, N.F. Azman, K. Chandrasekhar, M.S. Kalil, from wet biomass steam gasification with CaO/MgO, Int. J. Hydrogen Energy 40
A comprehensive review of microbial electrolysis cells (MEC) reactor designs and (2015) 8816–8823, https://doi.org/10.1016/j.ijhydene.2015.05.075.
configurations for sustainable hydrogen gas production, Alexandria Eng. J. 55 [127] L. Wei, S. Xu, L. Zhang, C. Liu, H. Zhu, S. Liu, Steam gasification of biomass for
(2016) 427–443, https://doi.org/10.1016/j.aej.2015.10.008. hydrogen-rich gas in a free-fall reactor, Int. J. Hydrogen Energy 32 (2007) 24–31,
[104] R. Tunold, A.T. Marshall, E. Rasten, M. Tsypkin, L.E. Owe, S. Sunde, Materials for https://doi.org/10.1016/j.ijhydene.2006.06.002.
electrocatalysis of oxygen evolution process in PEM water electrolysis cells, ECS [128] X. Ma, X. Zhao, J. Gu, J. Shi, Co-gasification of coal and biomass blends using
Trans 25 (2010) 103–117, https://doi.org/10.1149/1.3328515. dolomite and olivine as catalysts, Renew. Energy 132 (2019) 509–514, https://
[105] S. Bajracharya, M. Sharma, G. Mohanakrishna, X. Dominguez Benneton, D.P.B.T. doi.org/10.1016/j.renene.2018.07.077.
B. Strik, P.M. Sarma, D. Pant, An overview on emerging bioelectrochemical

14
T. Lepage et al. Biomass and Bioenergy 144 (2021) 105920

[129] F. Guo, X. Li, Y. Liu, K. Peng, C. Guo, Z. Rao, Catalytic cracking of biomass [152] G. Yang, J. Wang, Ultrasound combined with dilute acid pretreatment of grass for
pyrolysis tar over char-supported catalysts, Energy Convers. Manag. 167 (2018) improvement of fermentative hydrogen production, Bioresour. Technol. 275
81–90, https://doi.org/10.1016/j.enconman.2018.04.094. (2019) 10–18, https://doi.org/10.1016/j.biortech.2018.12.013.
[130] E. Yaghoubi, Q. Xiong, M.H. Doranehgard, M.M. Yeganeh, G. Shahriari, [153] K.S.K. Ismail, G. Najafpour, H. Younesi, A.R. Mohamed, A.H. Kamaruddin,
M. Bidabadi, The effect of different operational parameters on hydrogen rich Biological hydrogen production from CO: bioreactor performance, Biochem. Eng.
syngas production from biomass gasification in a dual fluidized bed gasifier, J. 39 (2008) 468–477, https://doi.org/10.1016/j.bej.2007.11.003.
Chem. Eng. Process. Process Intensif. 126 (2018) 210–221, https://doi.org/ [154] X. Wang, Y. Fang, Y. Wang, J. Hu, A. Zhang, X. Ma, H. Yang, L. Guo, Single-stage
10.1016/j.cep.2018.03.005. photo-fermentative hydrogen production from hydrolyzed straw biomass using
[131] H. Liu, T. Chen, D. Chang, D. Chen, D. Kong, X. Zou, R.L. Frost, Effect of Rhodobacter sphaeroides, Int. J. Hydrogen Energy 43 (2018) 13810–13820,
preparation method of palygorskite-supported Fe and Ni catalysts on catalytic https://doi.org/10.1016/j.ijhydene.2018.01.057.
cracking of biomass tar, Chem. Eng. J. 188 (2012) 108–112, https://doi.org/ [155] J. Cai, Y. Zhao, J. Fan, F. Li, C. Feng, Y. Guan, R. Wang, N. Tang, Photosynthetic
10.1016/j.cej.2012.01.109. bacteria improved hydrogen yield of combined dark- and photo-fermentation,
[132] A.S. Al-Rahbi, P.T. Williams, Hydrogen-rich syngas production and tar removal J. Biotechnol. 302 (2019) 18–25, https://doi.org/10.1016/j.jbiotec.2019.06.298.
from biomass gasification using sacrificial tyre pyrolysis char, Appl. Energy 190 [156] L. Yang, W. Liu, Z. Zhang, X. Du, J. Gong, L. Dong, Y. Deng, Hydrogen evolution
(2017) 501–509, https://doi.org/10.1016/j.apenergy.2016.12.099. from native biomass with Fe3+/Fe2+ redox couple catalyzed electrolysis,
[133] D. Yao, Q. Hu, D. Wang, H. Yang, C. Wu, X. Wang, H. Chen, Hydrogen production Electrochim. Acta 246 (2017) 1163–1173, https://doi.org/10.1016/j.
from biomass gasification using biochar as a catalyst/support, Bioresour. Technol. electacta.2017.06.124.
216 (2016) 159–164, https://doi.org/10.1016/j.biortech.2016.05.011. [157] T. Hibino, K. Kobayashi, M. Ito, Q. Ma, M. Nagao, M. Fukui, S. Teranishi, Efficient
[134] T. Miyazawa, T. Kimura, J. Nishikawa, K. Kunimori, K. Tomishige, Catalytic hydrogen production by direct electrolysis of waste biomass at intermediate
properties of Rh/CeO2/SiO2 for synthesis gas production from biomass by temperatures, ACS Sustain. Chem. Eng. 6 (2018) 9360–9368, https://doi.org/
catalytic partial oxidation of tar, Sci. Technol. Adv. Mater. 6 (2005) 604–614, 10.1021/acssuschemeng.8b01701.
https://doi.org/10.1016/j.stam.2005.05.019. [158] D. Unlu, N.D. Hilmioglu, Application of aspen plus to renewable hydrogen
[135] M. Ma, M. Müller, Investigation of various catalysts for partial oxidation of tar production from glycerol by steam reforming, Int. J. Hydrogen Energy 45 (5)
from biomass gasification, Appl. Catal. Gen. 493 (2015) 121–128, https://doi. (2020) 3509–3515. https://doi.org/10.1016/j.ijhydene.2019.02.106.
org/10.1016/j.apcata.2015.01.012. [159] Z. Liu, Economic analysis of energy production from coal/biomass upgrading;
[136] D.H. Kim, S.H. Kim, J.Y. Byun, A microreactor with metallic catalyst support for Part 1: hydrogen production, Energy Sources, Part B Econ, Plan. Policy. 13 (2018)
hydrogen production by partial oxidation of dimethyl ether, Chem. Eng. J. 280 132–136, https://doi.org/10.1080/15567249.2017.1410592.
(2015) 468–474, https://doi.org/10.1016/j.cej.2015.06.038. [160] D. Klein-Marcuschamer, H.W. Blanch, Renewable fuels from biomass: technical
[137] W. Żukowski, G. Berkowicz, Hydrogen production through the partial oxidation hurdles and economic assessment of biological routes, AIChE J. 61 (2015)
of methanol using N2O in a fluidised bed of an iron-chromium catalyst, Int. J. 2689–2701, https://doi.org/10.1002/aic.14755.
Hydrogen Energy 42 (2017) 28247–28253, https://doi.org/10.1016/j. [161] G. Collodi, G. Azzaro, N. Ferrari, S. Santos, Techno-economic evaluation of
ijhydene.2017.09.135. deploying CCS in SMR based merchant H2 production with NG as feedstock and
[138] G. Rabenstein, V. Hacker, Hydrogen for fuel cells from ethanol by steam- fuel, Energy Procedia 114 (2017) 2690–2712, https://doi.org/10.1016/j.
reforming, partial-oxidation and combined auto-thermal reforming: a egypro.2017.03.1533.
thermodynamic analysis, J. Power Sources 185 (2008) 1293–1304, https://doi. [162] A. Valente, D. Iribarren, J.L. Gálvez-Martos, J. Dufour, Robust eco-efficiency
org/10.1016/j.jpowsour.2008.08.010. assessment of hydrogen from biomass gasification as an alternative to
[139] S. Li, L. Guo, Stability and activity of a co-precipitated Mg promoted Ni/Al2O3 conventional hydrogen: a life-cycle study with and without external costs, Sci.
catalyst for supercritical water gasification of biomass, Int. J. Hydrogen Energy 44 Total Environ. 650 (2019) 1465–1475, https://doi.org/10.1016/j.
(2019) 15842–15852, https://doi.org/10.1016/j.ijhydene.2018.08.205. scitotenv.2018.09.089.
[140] S. Nanda, A.K. Dalai, J.A. Kozinski, Supercritical water gasification of timothy [163] I. Hannula, Hydrogen enhancement potential of synthetic biofuels manufacture in
grass as an energy crop in the presence of alkali carbonate and hydroxide the European context: a techno-economic assessment, Energy 104 (2016)
catalysts, Biomass Bioenergy 95 (2016) 378–387, https://doi.org/10.1016/j. 199–212, https://doi.org/10.1016/j.energy.2016.03.119.
biombioe.2016.05.023. [164] C. Acar, I. Dincer, Comparative assessment of hydrogen production methods from
[141] Q.M.K. Waheed, C. Wu, P.T. Williams, Pyrolysis/reforming of rice husks with a renewable and non-renewable sources, Int. J. Hydrogen Energy 39 (2014) 1–12,
Ni–dolomite catalyst: influence of process conditions on syngas and hydrogen https://doi.org/10.1016/j.ijhydene.2013.10.060.
yield, J. Energy Inst. 89 (2016) 657–667, https://doi.org/10.1016/j. [165] D. Graf, N. Monnerie, M. Roeb, M. Schmitz, C. Sattler, Economic comparison of
joei.2015.05.006. solar hydrogen generation by means of thermochemical cycles and electrolysis,
[142] G. Chen, J. Andries, H. Spliethoff, Catalytic pyrolysis of biomass for hydrogen rich Int. J. Hydrogen Energy 33 (2008) 4511–4519, https://doi.org/10.1016/j.
fuel gas production, Energy Convers. Manag. 44 (2003) 2289–2296, https://doi. ijhydene.2008.05.086.
org/10.1016/S0196-8904(02)00254-6. [166] S. Zhang, Prediction of selling price of hydrogen produced from methanol steam
[143] K. Akubo, M.A. Nahil, P.T. Williams, Pyrolysis-catalytic steam reforming of reforming, Energy Sources B Energy Econ. Plann. 13 (2018) 28–32, https://doi.
agricultural biomass wastes and biomass components for production of org/10.1080/15567249.2017.1402101.
hydrogen/syngas, J. Energy Inst. 92 (6) (2019) 1987–1996. https://doi.org/10.10 [167] Y. Zhang, T.R. Brown, G. Hu, R.C. Brown, Techno-economic analysis of
16/j.joei.2018.10.013. monosaccharide production via fast pyrolysis of lignocellulose, Bioresour.
[144] F. Chen, C. Wu, L. Dong, A. Vassallo, P.T. Williams, J. Huang, Characteristics and Technol. 127 (2013) 358–365, https://doi.org/10.1016/j.biortech.2012.09.070.
catalytic properties of Ni/CaAlOx catalyst for hydrogen-enriched syngas [168] K. Özdenkçi, C. De Blasio, G. Sarwar, K. Melin, J. Koskinen, V. Alopaeus, Techno-
production from pyrolysis-steam reforming of biomass sawdust, Appl. Catal. B economic feasibility of supercritical water gasification of black liquor, Energy 189
Environ. 183 (2016) 168–175, https://doi.org/10.1016/j.apcatb.2015.10.028. (2019) 116285. https://doi.org/10.1016/j.energy.2019.116284.
[145] F. Jin, H. Sun, C. Wu, H. Ling, Y. Jiang, P.T. Williams, J. Huang, Effect of calcium [169] D.A. Sladkovskiy, L.I. Godina, K.V. Semikin, E.V. Sladkovskaya, D.A. Smirnova, D.
addition on Mg-AlOx supported Ni catalysts for hydrogen production from Y. Murzin, Process design and techno-economical analysis of hydrogen
pyrolysis-gasification of biomass, Catal. Today 309 (2018) 2–10, https://doi.org/ production by aqueous phase reforming of sorbitol, Chem. Eng. Res. Des. 134
10.1016/j.cattod.2018.01.004. (2018) 104–116, https://doi.org/10.1016/j.cherd.2018.03.041.
[146] L. Dong, C. Wu, H. Ling, J. Shi, P.T. Williams, J. Huang, Promoting hydrogen [170] B. Pratto, M. Suzana, A.A. Longati, R.D.S. Júnior, A. José, G. Cruz, Experimental
production and minimizing catalyst deactivation from the pyrolysis-catalytic optimization and techno-economic analysis of bioethanol production by
steam reforming of biomass on nanosized NiZnAlOx catalysts, Fuel 188 (2017) simultaneous saccharification and fermentation process using sugarcane straw,
610–620, https://doi.org/10.1016/j.fuel.2016.10.072. Bioresour. Technol. 297 (2020) 122494. https://doi.org/10.1016/j.biortech.201
[147] Z.M.A. Bundhoo, Potential of bio-hydrogen production from dark fermentation of 9.122494.
crop residues: a review, Int. J. Hydrogen Energy 44 (2019) 17346–17362, [171] T. Sinigaglia, F. Lewiski, M.E. Santos Martins, J.C. Mairesse Siluk, Production,
https://doi.org/10.1016/j.ijhydene.2018.11.098. storage, fuel stations of hydrogen and its utilization in automotive applications-a
[148] S. Lunprom, O. Phanduang, A. Salakkam, Q. Liao, A. Reungsang, A sequential review, Int. J. Hydrogen Energy 42 (2017) 24597–24611, https://doi.org/
process of anaerobic solid-state fermentation followed by dark fermentation for 10.1016/j.ijhydene.2017.08.063.
bio-hydrogen production from Chlorella sp, Int. J. Hydrogen Energy 44 (2019) [172] K.Y. Show, Y. Yan, M. Ling, G. Ye, T. Li, D.J. Lee, Hydrogen production from algal
3306–3316, https://doi.org/10.1016/j.ijhydene.2018.06.012. biomass – advances, challenges and prospects, Bioresour. Technol. 257 (2018)
[149] T.T. Giang, S. Lunprom, Q. Liao, A. Reungsang, A. Salakkam, Enhancing hydrogen 290–300, https://doi.org/10.1016/j.biortech.2018.02.105.
production from Chlorella sp. Biomass by pre-hydrolysis with simultaneous [173] R.R. Gonzales, G. Kumar, P. Sivagurunathan, S.H. Kim, Enhancement of hydrogen
saccharification and fermentation (PSSF), Energies 12 (2019) 1–14, https://doi. production by optimization of pH adjustment and separation conditions following
org/10.3390/en12050908. dilute acid pretreatment of lignocellulosic biomass, Int. J. Hydrogen Energy 42
[150] J. Wang, Y. Yin, Fermentative hydrogen production using various biomass-based (2017) 27502–27511, https://doi.org/10.1016/j.ijhydene.2017.05.021.
materials as feedstock, Renew. Sustain. Energy Rev. 92 (2018) 284–306, https:// [174] M. Felgenhauer, T. Hamacher, State-of-the-art of commercial electrolyzers and
doi.org/10.1016/j.rser.2018.04.033. on-site hydrogen generation for logistic vehicles in South Carolina, Int. J.
[151] H.M. Zabed, S. Akter, J. Yun, G. Zhang, F.N. Awad, X. Qi, J.N. Sahu, Recent Hydrogen Energy 40 (2015) 2084–2090, https://doi.org/10.1016/j.
advances in biological pretreatment of microalgae and lignocellulosic biomass for ijhydene.2014.12.043.
biofuel production, Renew. Sustain. Energy Rev. 105 (2019) 105–128, https:// [175] M.F. Andrea, R.H. Sara, D.Z. Luca, S.S. Giovanni, B. Enrico, Techno-economic
doi.org/10.1016/j.rser.2019.01.048. analysis of in-situ production by electrolysis, biomass gasification and delivery
systems for Hydrogen Refuelling Stations: rome case study, Energy Procedia 148
(2018) 82–89, https://doi.org/10.1016/j.egypro.2018.08.033.

15
T. Lepage et al. Biomass and Bioenergy 144 (2021) 105920

[176] J. Proost, State-of-the art CAPEX data for water electrolysers, and their impact on emission characteristics of an adapted commercial four-cylinder spark ignition
renewable hydrogen price settings, Int. J. Hydrogen Energy 44 (2019) engine running on hydrogen – methane mixtures, Appl. Energy 113 (2014)
4406–4413, https://doi.org/10.1016/j.ijhydene.2018.07.164. 1068–1076, https://doi.org/10.1016/j.apenergy.2013.08.063.
[177] S. Schemme, J.L. Breuer, M. Köller, M. Sven, F. Walman, R.C. Samsun, R. Peters, [184] M. Talibi, P. Hellier, N. Ladommatos, ScienceDirect the effect of varying EGR and
D. Solten, H 2 -based synthetic fuels : a techno-economic comparison of alcohol , intake air boost on hydrogen-diesel co-combustion in CI engines, Int. J. Hydrogen
ether and hydrocarbon production, Hydrog. Ernergy. 5 (2019) 5395–5414, Energy 42 (2016) 6369–6383, https://doi.org/10.1016/j.ijhydene.2016.11.207.
https://doi.org/10.1016/j.ijhydene.2019.05.028. [185] N. Cihat Onat, M. Kucukvar, O. Tatari, Conventional , hybrid , plug-in hybrid or
[178] P. Rajangam, Hydrogen Fuel Cells as Green Energy, Hydrog. Fuel Cells as Green electric vehicles ? State-based comparative carbon and energy footprint analysis
Energy. Yang, P. (Eds.), Cases Green Energy Sustain. Dev. (Pp. 291-323). IGI Glob. in the United States, Appl. Energy 150 (2015) 36–49, https://doi.org/10.1016/j.
(n.d.) 291–293. https://doi.org/10.4018/978-1-5225-8559-6.ch011. apenergy.2015.04.001.
[179] L.M. Das, in: Hydrogen-fueled Internal Combustion Engines, Elsevier Ltd., 2016, [186] G. Wu, A. Inderbitzin, C. Bening, Total cost of ownership of electric vehicles
https://doi.org/10.1016/B978-1-78242-363-8.00007-4. compared to conventional vehicles : a probabilistic analysis and projection across
[180] M.A. Ahmed, Hydrogen fueled internal combustion Engine : a review, Int. J. market segments, Energy Pol. 80 (2015) 196–214, https://doi.org/10.1016/j.
Innov. Technol. Res. 4 (2016) 3193–3198. enpol.2015.02.004.
[181] W.W. Pulkrabek, Engineering fundamentals of the internal combustion engine, [187] H.S. Das, C.W. Tan, A.H.M. Yatim, Fuel cell hybrid electric vehicles : a review on
J. Eng. Gas Turbines Power 126 (2004) 126–198. power conditioning units and topologies, Renew. Sustain. Energy Rev. 76 (2017)
[182] C.M. White, R.R. Steeper, A.E. Lutz, The hydrogen-fueled internal combustion 268–291, https://doi.org/10.1016/j.rser.2017.03.056.
engine : a technical review, Int. J. Hydrogen Energy 31 (2006) 1292–1305, [188] E.A. Nanaki, C.J. Koroneos, Comparative economic and environmental analysis of
https://doi.org/10.1016/j.ijhydene.2005.12.001. conventional , hybrid and electric vehicles e the case study of Greece, J. Clean.
[183] P.M. Diéguez, J.C. Urroz, S. Marcelino-Sádaba, A. Pérez-Ezcurdia, M. Benito- Prod. 53 (2013) 261–266, https://doi.org/10.1016/j.jclepro.2013.04.010.
Amurrio, D. Sáinz, L.M. Gandía, Experimental study of the performance and

16

You might also like