Download as pdf or txt
Download as pdf or txt
You are on page 1of 25

Chemical Engineering Journal 452 (2023) 138980

Contents lists available at ScienceDirect

Chemical Engineering Journal


journal homepage: www.elsevier.com/locate/cej

Review

Photoreforming lignocellulosic biomass for hydrogen production:


Optimized design of photocatalyst and photocatalytic system
Cai Shi a, 1, Fuyan Kang a, 1, Yeling Zhu c, 1, Min Teng a, Junming Shi a, Houjuan Qi a,
Zhanhua Huang a, *, Chuanling Si b, *, Feng Jiang c, *, Jinguang Hu d, *
a
Key Laboratory of Bio-based Material Science and Technology, Ministry of Education, Material Science and Engineering College, Northeast Forestry University, Harbin
150040, China
b
Tianjin Key Laboratory of Pulp and Paper, College of Light Industry and Engineering, Tianjin University of Science and Technology, Tianjin 300457, China
c
Sustainable Functional Biomaterials Lab, Department of Wood Science, University of British Columbia, 2424 Main Mall, Vancouver V6T 1Z4, Canada
d
Department of Chemical and Petroleum Engineering, University of Calgary, 2500 University Drive, NW, Calgary, Alberta T2N 1N4, Canada

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

Keywords: Using renewable solar energy for clean hydrogen (H2) fuel production via photocatalysis is a promising low
Photocatalytic reforming carbon intensity strategy to address the energy crisis and environmental pollution issues. Biomass photorefinery
Biomass photorefinery is an emerging technology platform where biomass or its derivatives are employed as hole (h+) scavengers during
H2 evolution
photocatalysis, which not only reduces the thermodynamic barrier of the energy-demanding oxygen evolution
Solar energy
Biochemicals
reaction (OER, ΔE⁰ = − 1.23 V) in water splitting to favor H2 generation, but also endows the potential of biomass
valorization. This review provides an in-depth comprehensive overview of the co-production of H2 and value-
added biochemicals via lignocellulosic biomass photoreforming, with the special focuses on catalyst’s struc­
ture and photocatalytic system design. The photocatalyst structure is mainly elaborated from its band gap,
morphology and size, crystal phase, cocatalyst and surface group modification. Among them, the band gap
structure is crucial for the regulation of catalyst redox ability, which determines the high selectivity of photo­
catalyst to generate the desired biochemicals. As well as the catalytic system such as pH, solvent, the concen­
tration of catalyst and substrates, and the reaction temperature, matter much to enhance the solubility of
lignocellulose and the accessibility of catalysts to substrates. The performance of using raw and pretreated
biomass as h+ scavengers to boost H2 production as well as insights into their valorization pathways are deeply
discussed. Future perspectives and challenges for the further advancement of biomass photoreforming tech­
nologies are given at the end of this review.

1. Introduction effective technological alternative is to employ biomass or its derivatives


as reductive substrate to substitute OER for holes (h+) consumption,
Energy crisis is as one of the world’s most pressing issues that urges namely photocatalytic biomass reforming (or biomass photoreforming).
more efficient production of clean energy [1–3]. Sustainable H2, that can In this concept, upon photoexcitation, a photocatalyst generates h+ to
be produced through redox splitting of water via solar driven photo­ oxidize biomass and use the resultant electrons (e-) to reduce aqueous
catalysis, has been identified as a promising alternative to the dwindling proton (H+) to H2. In contrast to OER, the overall biomass photo­
fossil fuels [4–7]. Nevertheless, the conventional photocatalytic H2 reforming reaction is almost energy neutral, it can use highly abundant
production technologies are still challenged by the low quantum yields low-energy photons (visible and IR light), it also has the potential to
of water splitting, which is due to the large thermodynamic barrier of generate value-added by-products from biomass [10–13]. Ever since the
oxygen evolution reaction (OER, ΔE⁰ =− 1.23 V), rapid recombination of first report of using biomass as photoreforming substrate in 1980 s [14],
photogenerated charge carriers, and the existence of reverse-reactions lignocellulose, the most abundant biomass on earth, has demonstrated a
[8,9]. To promote photocatalytic water splitting for H2 production, an great potential biomass photoreforming. The reason for that is,

* Corresponding authors.
E-mail addresses: huangzh1975@163.com (Z. Huang), sichli@tust.edu.cn (C. Si), feng.jiang@ubc.ca (F. Jiang), jinguang.hu@ucalgary.ca (J. Hu).
1
These authors contributed equally.

https://doi.org/10.1016/j.cej.2022.138980
Received 20 June 2022; Received in revised form 23 August 2022; Accepted 30 August 2022
Available online 6 September 2022
1385-8947/© 2022 Elsevier B.V. All rights reserved.
C. Shi et al. Chemical Engineering Journal 452 (2023) 138980

lignocellulose not only serves as an e- donor and/or h+ scavenger to design of photocatalyst’s structure and catalytic system on photo­
promote the separation and consumption of the photogenerated reforming of different type of biomass are discussed in detail, starting
charges, but also be simultaneously transformed into various value- from the three major components of lignocellulose (cellulose, hemicel­
added biochemicals. Meanwhile, the photocatalytic H2 evolution effi­ lulose, and lignin), followed by the raw biomass, wastes, and the small
ciency is enhanced as a result of being coupled with the photocatalytic molecules derived from biomass. Our opinions on the perspectives and
lignocellulose oxidation [15–17]. future challenges toward photocatalytic biomass reforming for H2 evo­
Despite these advantages, the incumbent lignocellulose photo­ lution are given for the further advancement of novel photoreforming
reforming technology is still incapable of meeting the industrial expec­ technologies.
tations. One bottleneck is the intractable deconstruction of common
lignocellulosic feedstocks, primarily due to the sophisticated structure of 2. Fundamental principles of biomass photoreforming
lignocellulose that has evolved to protect against microbial and photo
degradation in nature [18,19]. Cellulose (a semi-crystalline poly­ As shown in Fig. 1, biomass photoreforming reaction is driven by the
saccharide, >40 % in dry weight), hemicelluloses (branched poly­ excited state of semiconductor under solar irradiation, which consists of
saccharides, 20–40 %), and lignin (crosslinked phenolic polymer, less two separate half-reactions, namely H2 evolution reaction (HER) and
than35 %) represent the three largest fractions of lignocellulose. The biomass oxidation reaction (BOR). In general, photoreforming reaction
strong inter-chains hydrogen bonding within the cellulose microfibrils is mainly accomplished by three sub-processes: 1) activation of the
and the encrusting recalcitrant lignin molecules make it resistant to photocatalyst to generate photoexcited h+-e- pairs under the irradiation
chemical transformation [20,21]. (Fig. 1) Other drawbacks encountered of incident light with energy greater than the photocatalyst band gap
in photoreforming lignocellulose include low selectivity toward energy. The e- present in valence band (VB) can be excited and trans­
partially oxidized products (especially with water as the solvent), the ferred to the conduction band (CB) of the semiconductor, while the same
inhibitory effects of by-products in the liquid phase, and the demand of number of h+ are simultaneously created and remain in the VB; 2)
high-energy UV light to activate certain photocatalysts [22–24]. migration of both types of the photogenerated carriers to the photo­
To achieve high-efficiency photoreforming of lignocellulose and its catalyst surface; 3) utilization of the photogenerated carriers by their
derivatives for H2 production and biomass valorization, considerable corresponding substrates to produce H2 and biochemicals, i.e., e- on the
progress have been witnessed in the past years in designing photo­ photocatalyst surface reduces protons to H2, while the h+ oxidizes
catalysts with optimized structure and/or developing novel photo­ biomass-derived reactants into corresponding chemicals.
catalytic system. Currently, there are several excellent reviews just focus HER is usually substrate independent, and is primarily a process in
on the photocatalysts or systems designs for the photocatalytic H2 pro­ which the protons from water are reduced by photogenerated e-.
duction from H2O [25] or photoreforming of organics for oxidation re­ Biomass photoreforming has a much smaller thermodynamic barrier
actions [26–28]. Afterwards, Qi et al. summarized the cooperative than that of photodecomposition of water, making the H2 evolution
coupling of oxidative organic synthesis (from alcohol, ether, hydrocar­ reaction more likely to occur in theory. In addition, it facilitates more
bon, and amine, etc. to value-added chemicals) and H2 production over sufficient use of solar energy with a wider range of wavelengths (e.g.
semiconductor-based photocatalysts [29]. In 2021, our group published visible or even IR light), which is vital for practical application [11]. The
a mini-review on photoreforming biomass for the production of H2 and photoreforming HER reaction carried out in D2O clearly indicates that
chemicals. However, the effect of photocatalysts and systems optimi­ the generated H2 is mainly derived from water rather than biomass [21].
zation on the generation of H2 and chemicals in photoreforming are not For some specific reactions, like the photoreforming process of benzyl
particularly clear and complete [30]. To further deepen the under­ alcohol, furfural alcohol, methanol, and ethanol, the protons could be
standing of the effects of catalyst structure and catalytic system opti­ from alcohols substrates [31]. Yet biomass as proton sources for H2
mization on lignocellulosic biomass photoreforming for H2 and evolution remains to be further explored.
chemicals production, this review will provide a comprehensive over­ BOR is a more complex process that involves a multi-step conversion
view on it. Furthermore, the performance of using lignocellulose and its of biomass substrates. Different kinds of biomass and reaction systems
derivatives as h+ scavengers as well as insights into their valorization including catalysts and solvent types, pH and so on will affect the re­
pathways are deeply discussed. Specifically, the effects of the optimized action pathway and mechanism of biomass in photoreforming process.

Fig. 1. Chemical structures of lignocellulosic components and photoreforming biomass to evolve H2 and oxidized organics on semiconductor.

2
C. Shi et al. Chemical Engineering Journal 452 (2023) 138980

Briefly, biomass consumes h+, ⋅OH, ⋅O–2 or 1O2 who can be transferred acid-base sites (Summarized in Table 1).
directly to the chemisorbed biomass substrates [32,33] to generate
corresponding chemicals in the process of photoreforming, in which
3.1. Photocatalyst band gap structure
active species were involved mainly depends on the oxidation capacity
determined by the band gap of the photocatalyst itself. In general, the
As described in Section 2, photoreforming reaction is mainly
consumption of oxidized active species would make more available e- for
accomplished by three sub-processes, and the band structure of photo­
HER, which is conducive to the improvement of HER efficiency.
catalyst mainly affects Steps 1 and 2. To ensure the forward reaction of
Nevertheless, the conversion pathway of biomass is up to the specific
photocatalytic HER during water splitting, the band gap of photocatalyst
reaction system. Taking the simple monosaccharide-glucose as an
needs to be greater than the potential barrier of HER (1.23 eV).
example, it could be photoreformed to generate small molecular acids
Considering the thermodynamic loss and overpotential in the reaction
such as lactic acid by isomerization, retro-aldol reaction, and dehydra­
process, the band gap energy in practical application typically needs to
tion [34]; also the C–C bond in glucose could be cleaved selectively,
be>1.8 eV [11]. However, photocatalyst with high band gap energy
and the glucose can be gradually converted into arabinose, erythrose,
suffers from insufficient utilization of solar energy. For instance, to make
and formic acid in photocatalytic glucose reforming [35]. This shows the
full use of the visible light that accounts for approximately 46 % of the
liquid products differ immensely with diverse photoreforming systems.
total solar energy [37], the band gap of the photocatalyst is preferably
It should be pointed out that the relationship between HER and BOR is
less than 3.2 eV. Therefore, the theoretical optimal band gap energy of
relatively complicated, which will be analyzed in detail in the following
the material used for photocatalytic HER should be between 1.8 and 3.2
sections combined with specific reaction systems.
eV [38–40].The absorption characteristics of a material depend on its
bandgap value. So, the selection of light source should be consistent with
3. Influences of photocatalyst structure on photoreforming H2
the absorption characteristics of the semiconductor in practical photo­
generation
reforming process. Other than the band gap energy, the band gap dis­
tribution is equally important for HER. The CB potential of photocatalyst
The structure of photocatalyst plays an important role in determining
needs to be more negative than the reduction potential of H2 (E0 = -0.41
the HER activity by affecting the physicochemical and optical proper­
V vs standard hydrogen electrode, at pH = 7), and the VB potential
ties, as well as the separation efficiency of the photogenerated charges
should be more positive than the oxidation potential of the sacrificial
[36]. In this part, the effects of catalyst structure on biomass photo­
agent (i.e., the biomass). It is worth mentioning that the light with less
reforming H2 production will be discussed from the following four as­
energy and longer wavelength (such as visible and IR light) in solar
pects: 1) band gap structure; 2) morphology and size; 3) crystal phase;
energy that used to be less useful in pure water splitting can be more
and 4) other factors such as cocatalyst loading, modification of transi­
sufficiently utilized to activate H2 production reaction during biomass
tion metals elements for selective adsorption of substrates and surface
phototreforming [11].

Table 1
Influence of photocatalyst structure on photoreforming H2 production performance.
Influence factor Entry Photocatalyst Substrate Light source, intensity (mW H2 evolution rate (mmol g- AQY/% (λex/ Reference
(Concentration, g L-1) cm− 2) 1 − 1
h ) nm)

Band gap structure 1 TiO2-NiO (0.50) Lignin UV–Vis 23.52 6.46 (3 6 5) [54]
TiO2 (0.50)NiO 6.60 2.54
(0.50) 2.11 (3 6 5)0.92
(3 6 5)
2 Sn0.24-Ru0.68/TiO2 (0.30) Glucose UV ~4.00 0.41 (3 6 5) [55]
Ru0.68/TiO2 (0.30) 0.33 0.30
Sn/TiO2 (0.30) 0.06 (3 6 5)0.06
TiO2 (0.30) 0.06 (3 6 5)0.01
(3 6 5)
3 ZnxCd1− xS-P (0.20)ZnS-P 5-HMF LED 0.79 – [63]
(0.20)CdS-P 0.006
(0.20) 0.06
4 Zn0.6Cd0.4S homojunction (2.00) Glucose Xe lamp 0.69 ~0.18 (4 2 0) [34]
ZnS
(2.00)CdS 0.02
(2.00) 0.11
5 ZnS-g-C3N4 (1.00) Glucose Xe lamp 0.21 – [70]
g-C3N4 (1.00) 0.08
6 UCN-NA0.5 (0.67)UCN Glucose LED ~1.91 – [71]
(0.67) 0.16
Morphology and 7 CdS/CdOx (0.50 μM) Lignocellulose Solar 2.57 – [21]
size light simulator
8 CdS nanosheets (0.50)CdS Lactic acid Xe lamp ~3.75 – [79]
nanospheres ~6.25
(0.50)CdS nanorods ~6.25
(0.50)CdS nanowires ~5.38
(0.50)Commercial CdS ~5.13
(0.50)
Plasma effect 9 3DOM TiO2-Au (0.20) Glucose Xe lamp ~10.90 [45]
TiO2 (0.20) ~0.67
Element 10 Ni-CdS (1.00) Furfural 8 W LED ~2.59 – [42]
modification alcohol
5-HMF ~0.47 –
Acid-base sites 11 SO2-
4 -P25-NixSy (0.20) Cellulose Xe lamp 3.02 – [94]
SO2-
4 -P25(0.20) (0.20) 0.19 –
Pt/P25 (0.20) 0.04 –

3
C. Shi et al. Chemical Engineering Journal 452 (2023) 138980

Even though semiconductor with favored band gap energy and dis­ clearly indicating that lignin can be effectively depolymerized by the
tribution was selected, its photocatalytic efficiency can still be low due accelerated separation of the photogenerated carriers in TiO2-NiO het­
to the possible high recombination rate of the photogenerated carriers erojunction photocatalyst. At the same time, though the mechanism was
and the limited reactive sites. To overcome such limitation, it is neces­ not clear, the depolymerized lignin units can be partially converted into
sary to regulate the structure of the band gap by conducting element fatty acids to consume more oxidative species, furthermore, promoting
doping, heterojunction or homojunction construction, and surface the production of H2.
functional group modification. The effects of band gap structure of Gu and coworkers constructed two heterojunctions by grafting Ru/
commonly used photocatalysts on the biomass photoreforming will be TiO2 with Sn and tested them for H2 production by using different
reviewed in the following section. biomass small molecules (glucose, methanol, ethanol, and glycerol) as
photoreforming reactants [55] (Entry 2 in Table 1). Two types of het­
3.1.1. TiO2 based materials erojunctions were formed within the photocatalyst by Ru and Sn
Among various semiconductive materials, TiO2, CdS, CdS-based loading, as shown in Fig. 2e, which enabled the formation of a built-in
bimetallic sulfides, and g-C3N4 have been widely studied in photo­ electric field within the photocatalyst. In the photoreforming process,
reforming biomass for H2 production due to their advantages of suitable RuO2 functioned as h+-acceptor to facilitate the subsequent oxidation
band gap structure and low cost [41–44]. Among them, TiO2 is the most reaction of biomass substrates, and the grafted atomic Sn species served
extensively and deeply studied owing to its non-toxicity, high chemical as e--acceptor to produce H2. As a result, the photogenerated charges can
stability, and low cost. However, pure TiO2 exhibits limited efficiency in be effectively separated, leading to improved photoreforming activity.
biomass photoreforming HER, due to its large band gap energy (around
3.20 eV) and the high recombination rate of photogenerated carriers 3.1.2. CdS and CdS-based bimetallic sulfides
[45,46]. Both challenges can be overcome by modifying the band gap For CdS photocatalyst, its narrow band gap (about 2.30 eV) enables a
structure of TiO2 for greater H2 production [47,48]. wide photoresponse range and favors solar energy harvesting with wider
Constructing heterojunction is a common method to adjust the band spectrum, yet simultaneously leading to serious photocorrosion issues
gap of TiO2 for the improved photoreforming performance [49]. The so- [56]. To restrain photocorrosion for the improved photocatalytic per­
called heterojunction is the interface region formed by the contact of formance (including organic pollutants degradation, H2 production
two different semiconductors [50–53]. For example, Zhao and his col­ from water splitting and biomass photoreforming, etc.) [48,57,58], one
leagues prepared TiO2-NiO core–shell heterojunction by one-step hy­ practical approach is to adjust the band gap structure of CdS. Early
drothermal method, which was used for photocatalytic H2 production research has demonstrated the effectiveness of conducting element
with kraft lignin as substrate [54] (Entry 1 in Table 1). TiO2 and NiO doping or heterojunction (or homojunction) construction in achieving
were closely combined to construct built-in electric field by forming type these targets [59–62]. Ye et al. prepared P-doped ZnxCd1-xS (ZnxCd1-xS-
II heterojunction, which significantly accelerated the separation rate of P) by hydrothermal and high-temperature calcination [63] (Entry 3 in
e--h+ pairs and further improved the photocatalytic performance of TiO2 Table 1). Firstly, the photocatalytic H2 evolution from pure water
(Fig. 2a-d). The results showed that the H2 production rate of TiO2-NiO without the assistance of h+ sacrificial agent was evaluated by ZnxCd1-xS
core–shell catalyst reached 23.50 mmol g-1 h− 1, which was about 3.56 and ZnxCd1-xS-P (0 ≤ x ≤ 1) under visible light irradiation (Fig. 3a).
and 11.1 times higher than that of pristine TiO2 (6.60 mmol g-1 h− 1) and Regardless of the × value, no H2 was generated for all ZnxCd1-xS systems
NiO (2.11 mmol g-1 h− 1), respectively (Fig. 2a-b). Some fatty acids were in the absence of sacrificial reagents after 2 h irradiation, while all
also obtained in the liquid phase as the lignin photoreforming products. systems carrying ZnxCd1-xS-P exhibited relatively high photocatalytic
To further investigate the reason for improved photoreforming activity, activities under the same condition. Among all the tested, Zn0.5Cd0.5S-P
the structure of kraft lignin before and after the reaction was charac­ gave a maximum H2 production rate (0.42 mmol g-1 h− 1). This can be
terized by Hetero Single Quantum Coherence-Nuclear Magnetic Reso­ attributed to the variation of the band gap of ZnxCd1-xS after P doping
nance (HSQC NMR) by using guaiacyl units (G2) as the internal standard (from 2.49 eV for ZnxCd1-xS to 2.43 eV for ZnxCd1-xS-P), and the
(Fig. 2c-d). The results showed that the content of β-O-4 in kraft lignin endowed visible light activity of ZnxCd1-xS. When using the same
decreased significantly after 12 h reaction catalyzed by TiO2-NiO, optimal catalyst (Zn0.5Cd0.5S-P) for photoreforming

Fig. 2. Photocatalytic H2 production using heterojunction TiO2-NiO core–shell catalyst: (a) Photocatalytic H2 evolution activities in water-methanol system; (b)
Corresponding rates of pure NiO, pure TiO2 and different nanocomposites; HSQC NMR spectrum of (c) lignin control and (d) depolymerized lignin. Reprinted with
permission from ref. [54]. Copyright 2021 Elsevier. Photoreforming H2 production over sn-Ru/TiO2: (e) Proposed reaction pathways for the photocatalytic reforming
of biomass-derived molecule. Reprinted with permission from ref. [55], Copyright 2013 Elsevier.

4
C. Shi et al. Chemical Engineering Journal 452 (2023) 138980

Fig. 3. Photoreforming H2 production performance of ZnxCd1-xS-P (0 ≤ x ≤ 1): (a) Comparison of photocatalytic hydrogen evolution rates of ZnxCd1-xS-P (0 ≤ x ≤ 1)
solid solutions from pure water without using of any electron sacrificial agents; (b) Comparison of photocatalytic H2 evolution rates of Zn0.5Cd0.5S-P from pure water
and 5-HMF solution at various times; (c and d) GC–MS traces of photocatalytic oxidation of 5-HMF by Zn0.5Cd0.5S-P. Reprinted with permission from ref. [63],
Copyright 2018 Elsevier. Photoreforming H2 evolution activity over Zn1-xCdxS homojunction: (e) Schematic atomic model of a Zinc blende-Wurtzite superlattice
structure, the twin boundaries in homojunction, and the migration of charge carriers; (f) Bandgap structure of various Zn1-xCdxS solid solutions; (g) Photocatalytic
hydrogen evolution rates of Zn1-xCdxS solid solutions in glucose solution. Reprinted with permission from ref. [34], Copyright 2021 Elsevier.

hydroxymethylfurfural (5-HMF), a H2 evolution rate of up to 0.79 mmol However, the poor visible light response and high photogenerated
g-1 h− 1 was obtained. More interestingly, Zn0.5Cd0.5S-P can maintain charges recombination rate limit the biomass photoreforming perfor­
good photocatalytic stability when coupled with 5-HMF photocatalytic mance of carbon nitride. Similar to the case of TiO2 and CdS, the biomass
oxidation, as supported by a good retention of the H2 generation rate photoreforming activity of carbon nitride can be tuned by adjusting the
(0.62 mmol g-1 h− 1) after 8 h light irradiation (Fig. 3b). In addition, the band structure, following strategies such as heterojunction construction
liquid organic products from 5-HMF oxidization were further identified and functional groups modification [53,68–71]. Xu et al. prepared ZnS-
by GC–MS (Fig. 3c and d). The peaks at around 19.50 and 14.80 min can modified g-C3N4 (MCN, DCN, and UCN derived from melamine,
be ascribed to 5-HMF and 2, 5-diformylfuran (DFF), respectively, dicyandiamide, and urea) composites by dipping and calcination, with
showing clearly weakened 5-HMF peak and increased DFF peak over the the formed heterojunctions denoted as ZMCN, ZDCN, and ZUCN,
irradiation time. All these results suggested that P doping plays a respectively [70] (Entry 5 in Table 1). The photoreforming H2 produc­
favorable role on narrowing the band gap and improving the photo­ tion activity of the three heterojunctions was evaluated using glucose as
reforming activity of CdS photocatalyst. model substrate. Compared with bare g-C3N4 samples (MCN and DCN),
Beyond heterojunction, constructing homojunction is also a common the glucose photoreforming performances of the corresponding hetero­
method to adjust the band gap of CdS materials [23,50–52,64–66]. Zhao junctions ZMCN and ZDCN were significantly improved. In contrast,
et al. introduced ZnS and prepared ZnxCd1-xS homojunction to produce ZMCN achieved the highest H2 generation productivity (0.07 mmol g-1
H2 and photoreform glucose into lactic acid [34] (Entry 4 in Table 1). As h− 1), which was 2.51 times than that of MCN (0.03 mmol g-1 h− 1)
these two crystal structures have almost the same lattice constants, a (Fig. 4a). The results showed the high-efficiency heterojunction was
homojunction structure by bridging cubic zinc blende and hexagonal formed between ZnS and g-C3N4, which broadened the catalyst’s pho­
wurtzite segments can be constructed with long-range order, resulting in toresponse range and enhanced the separation of photocarriers (Fig. 4b-
a twinning superlattice. The formation of homojunction greatly pro­ c). Surprisingly, the efficiency of ZUCN decreased by 36.6 % as
moted the separation of photocharged carriers (Fig. 3e). A major compared to the pure UCN (Fig. 4a). This might be ascribed to the wider
advantage of such design is that by controlling the molar ratio of Zn/Cd, band gap of UCN (2.81 eV) and the associated CB position change, which
the band gap energy can be well regulated. As can be obviously seen was closer to CB of ZnS than those of MCN and DCN (Fig. 4b-d). These
from Fig. 3f, the band gap gradually widens with the increase of Zn results clearly demonstrated the impact of band gap structure on pho­
content in the ZnxCd1-xS solid solution, from 2.25 eV (CdS) to 3.55 eV toreforming H2 evolution.
(ZnS). When x = 0.6, a perfect homojunction is formed in Zn0.6Cd0.4S Introducing functional groups can also affect the band gap of g-C3N4.
(Fig. 3e), which exhibited the optimal H2 evolution rate of 0.69 mmol g-1 Liu et al. synthesized edge functionalized carbon nitride via an in-situ
h− 1. Both the pristine CdS and ZnS exhibited a much lower H2 genera­ C–N coupling method. The carbon nitride-N-acetylethanolamine
tion rate than that of Zn0.6Cd0.4S (Fig. 3g) [34], indicating that the as­ (UCN-NA) composites were synthesized by thermal pyrolysis of additive
sembly of homojunction structure indeed affected the band gap of NA and precursors urea, in which the hydroxyethyl groups were grafted
photocatalysts and improved the separation efficiency of photo­ to the heptazine rings by substituting its terminal amino groups [71]
generated charge carriers. (Entry 6 in Table 1). The band gap structures, electronic distribution,
and energy-level structures were elucidated by density functional theory
3.1.3. g-C3N4 based materials (DFT) calculation and Gaussian 09 (Fig. 4e-h). The results showed that
Carbon nitride is the most representative one among non-metallic the introduction of the hydroxyethyl created discrete energy levels in
photocatalysts. It has received extensive attention in the field of the bandgap and enhanced the absorption of visible light (Fig. 4e-f);
biomass photoreforming HER due to its environmentally friendly char­ Also, redistribution of e- clouds occur, and the spatial separation of
acteristic, low cost, and unique two-dimensional structure [67]. LUMO and HOMO in UCN-NA might also favor the separation of h+-e-

5
C. Shi et al. Chemical Engineering Journal 452 (2023) 138980

Fig. 4. g-C3N4/ZnS composites pho­


toreforming H2 generation form
monosaccharide: (a) Photoreforming
H2 evolution on pure g-C3N4 (MCN,
DCN, and UCN) and g-C3N4/ZnS
(ZMCN, ZDCN, and ZUCN); Sche­
matic band structures for (b) ZMCN,
(c) ZDCN, and (d) ZUCN. Reprinted
with permission from ref. [70],
Copyright 2019 Elsevier. Edge func­
tionalization of terminal amino group
in carbon nitride by in-situ C–N
coupling: Calculated band structure of
(e) UCN and (f) UCN-NA by using DFT
calculation. Chemical structures of
the polymeric models of (g) UCN and
(h) UCNNA together with the calcu­
lated HOMO and LUMO energy levels;
(i) Photocatalytic H2 production using
UCN and UCN-NA0.5 with 100 mg
xylose, glucose, fructose and arabi­
nose under white LED irradiation for
4 h. Reprinted with permission from
ref. [71], Copyright 2020 Elsevier.

pairs compared to UCN (Fig. 4g-h). The UCN-NA0.5 (the volume of NA swelling under strong alkaline conditions, but also the favorable
solution added is 0.50 mL) displayed the detectable photoreforming morphology and size of the photocatalysts. Quantum dot-like CdS with
HER activity with various monosaccharides as substrates (Fig. 4i). In small size has a large SSA and abundant surface-active sites (Fig. 5a-b),
detail, UCN-NA0.5 showed higher H2 production rates about 6 and 8.5 which can achieve effective contact between CdS photocatalyst and the
times than that of UCN for the photoreforming of glucose and xylose, dissolved cellulose molecules, ultimately resulting in a significantly
respectively. In conclusion, the significantly improved H2 evolution enhanced H2 production rate in the system. Even for the less-processed
activity further confirmed that the band gap structure of photocatalyst lignocellulosic substrates (wood from a tree branch, Fig. 5c), the pro­
truly plays a critical role in photoreforming reaction. posed quantum dot-like CdS catalyst still gave remarkable cellulose
Different from traditional photocatalytic HER, little attention has photoreforming HER activity, i.e., H2 production rate of over 5 mmol g-1
been paid to the development of the photocatalyst with suitable band h− 1. In another study, Song and his colleagues prepared CdS with
gap structure for biomass photoreforming process. In addition to satis­ different morphologies (nanospheres, nanorods, nanowires, and nano­
fying the reduction potential of HER, controlling the oxidation ability of sheets) by a simple hydrothermal method (Fig. 5d), and investigated
the VB potential is also crucial to improve the selectivity of biomass their photocatalytic performance for producing alanine (where glucose
conversion. Otherwise, excessive oxidation products of the substrate or and ammonia were used as biomass substrates and nitrogen source,
other unexpected liquid products may accumulate on the surface of the respectively) and H2 (where lactic acid was used as substrates) [79]
photocatalysts, resulting in its deactivation that is detrimental to its (Entry 8 in Table 1). The results indicated that the optimum photo­
long-term stability [72]. Meanwhile, more consideration should be forming H2 and alanine generation rates were achieved on catalysts with
given to the effects of the pathway and degree of biomass oxidation on different morphologies, respectively. CdS nanosheets exhibited the
H2 generation efficiency in photoreforming reaction. This is dramati­ worst H2 production ability but the optimal alanine generation, while
cally vital to promote the development of the overall biomass photo­ CdS nanorods showed the opposite (Fig. 5e-f). Free radicals capture
reforming reaction. experiments showed that only nanosheets could generate oxygen-centric
free radicals (⋅OCR), which may be the primary reason for the highest
alanine generation rate. This work supported that the active species
3.2. Photocatalyst morphology and size produced in the reaction process could be controlled by photocatalyst
morphology. Meanwhile, it is also demonstrated that the production rate
Morphology and size of photocatalysts are two important factors that of H2 and liquid product does not necessarily follow the same trend,
govern their photocatalytic performance. Firstly, morphology and size which requires case-by-case analysis in specific photoreforming
of photocatalysts will affect its physicochemical properties, including reaction.
specific surface area (SSA) and surface active sites, which are directly
correlated with the effective contact between the catalysts and sub­
strates and subsequently impact the photoreforming reaction [73–76]. 3.3. Photocatalyst crystal phase
In general, the SSA and active sites of materials tend to increase as the
particle size decreases. In some cases, photocatalyst’s morphology also The crystal phase of catalyst is another key factor in biomass pho­
affects the generation of reactive species, which plays a decisive role in tocatalytic reforming process, which has an impact on HER by affecting
the biomass photoreforming reaction [77,78]. Here, the influence of the selectivity to substrates or the acidity and alkalinity of catalyst
photocatalyst morphology and size on photoreforming HER activity is surface. It also affects the reaction pathway of biomass substrates and
reviewed. products types in the photoreforming process, yielding different high
Wakerley and his colleagues prepared CdS quantum dots by heat value-added chemicals [80].
reflow method for biomass photoreforming H2 evolution (Fig. 5a-c) [21]
(Entry 7 in Table 1). With cellulose as photoreforming substrates under 3.3.1. The selectivity of photocatalyst crystal phase to lignocellulose
strong alkaline conditions, up to 5.57 mmol g-1 h-1 H2 production rate substances
was achieved. Such excellent photoreforming HER performance is not In the traditional photocatalytic process, the difference in crystal
only attributed to the improved catalyst stability and enhanced cellulose type has no obvious effect on the oxidation pathway of common

6
C. Shi et al. Chemical Engineering Journal 452 (2023) 138980

Fig. 5. Photoreforming lignocellulose for H2 evolution


over CdS/CdOx: (a) TEM image of CdS/CdOx after
isolation from 10 M KOH; (b) Corresponding size dis­
tribution; (c) Photoreforming H2 generation from crude
lignocellulose when CdS/CdOx suspended in alkaline
solution and irradiated with sunlight. Reprinted with
permission from ref. [21], Copyright 2017 Nature.
Photoreforming biomass for synthesis alanine over
different morphology of CdS: (d) TEM image of a com­
mercial CdS, nanospheres, nanorods, and nanowires
(from left to right), scale bar = 200 nm; (e) In situ ESR
spectra for CdS nanosheets and nanorods in a mixture
solution of lactic acid and ammonia in the presence of
spin-trapping agent 5, 5-dimethyl-1-pyrroline N-oxide
(DMPO) with or without visible-light irradiation; (f) H2
production performance of CdS with different mor­
phologies. Reprinted with permission from ref. [79],
Copyright 2020 Nature.

sacrificial agents (triethanolamine, sodium sulfide, etc.). However, it 3.3.2. Surface acidity/basicity property
does play a decisive role in the transformation pathway of biomass and For the general photoreforming reaction, the desired H2 is produced
products type in the photoreforming oxidation half-reaction, which af­ by dehydrogenation of an oxygen-containing matrix, accompanied by
fects the corresponding HER rate as well [81]. Iervolino et al. prepared CO2 generation. Unfortunately, in certain systems the production of CO2
TiO2 co-doped with fluorine and Pt (Pt-F-TiO2) for glucose photo­ and H2 can be replaced by CO and H2O, making it less desirable for
reforming H2 evolution [81]. The results showed that the crystalline direct application of H2 to fuel cells. Further investigation elaborated
phase of TiO2 in Pt-F-TiO2 catalyst was precisely in anatase form. that such selectivity in reaction pathway is closely associated with the
Despite a lower selectivity on arabinose and erythrose compared to acidity/basicity of the photocatalyst used. For instance, formic acid-like
rutile, the anatase phase was more inclined to break the C–C bond in the species are a common type of intermediate chemicals in the photo­
glucose molecules adsorbed on catalyst surface. This promoted the reforming biomass process. It has been identified that such formic acid-
generation of H2, CO2, and gluconic acid, whereas the production of like species tend to transform to CO and H2O when in contact with the
arabinose and erythrose on Pt-F-TiO2 was inhibited. Under the optimum acid sites of photocatalysts, while the basic surfaces of catalysts are
conditions, the H2 yield of Pt-F-TiO2 can reach 0.59 mmol g− 1 h− 1. In favorable for their dehydrogenation to produce CO2 and H2 [47,82].
addition to the effect of photocatalyst crystal on the oxidation products The surface acidity/basicity of catalysts, as well as the ⋅OH groups
selectivity, it is noteworthy that the BOR and HER are interrelated, but present on their surface, are impacted by the ratio of different crystal
neither in a simple synergistic nor in an antagonistic way. Of course, nor phases in photocatalyst and hence the hydroxylation degree of adsorbed
can we primitively judge the oxidation pathway and the liquid product H2O molecules [82–85]. Xu et al. synthesized Pt/P25 photocatalysts
types based on the performance of photoreforming H2 evolution only. with different ratio of anatase and rutile phases, which were applied for
Therefore, it is significantly necessary to analyze the relationship be­ the photoreforming of methanol, glycerol, and glucose for H2 generation
tween BOR and HER in combination with the specific reaction system in [82]. The results showed that the ⋅OH group on the interface between
the biomass photoreforming process. the anatase and rutile phases was removed after heat treatment, which
resulted in an increase of surface basicity, leading to a favored reduction

7
C. Shi et al. Chemical Engineering Journal 452 (2023) 138980

in the CO/H2 ratio (the CO concentration in H2) of gas products. Under 2) Enhanced selective adsorption of photocatalysts toward substrates by
the optimized condition, the CO/H2 ratio obtained by the optimal TiO2- introducing transition metals; and 3) Modification of surface acid/base
based photocatalyst was less than 5 ppm. In addition, the formic acid sites. These practices are capable of promoting the separation of pho­
photoreforming performance of P25-x%R (x%R means the ratio of rutile tocarriers, leading to more active sites/species available for photo­
in P25) were investigated to study the underlying reason of significantly reforming reaction and therefore improved H2 evolution efficiency.
inhibited CO formation. The results suggested that the production
selectivity between CO and CO2, related to the dehydration or dehy­ 3.4.1. Cocatalysts
drogenation of formic acid, is governed by the acidity/basicity of the Cocatalysts, as an important component of photocatalysts, mainly
P25 surface; the acid sites are conducive to the dehydration reaction, contribute to the efficient e- transportation and carriers’ separation.
while the basic P25 surfaces are favorable for the dehydrogenation. Depositing noble metals, metal phosphides and sulfides is a promising
Thermal treatment of P25 is capable of increasing the photocatalyst’s method, which can effectively promote the photoreforming efficiency.
basicity by removing the ⋅OH at the anatase/rutile interface during the The enhancement of the photocatalytic activity of loaded noble
amorphous to crystalline phase and/or from anatase to rutile trans­ metal particles is mainly derived from the plasma effect. Firstly, the
formation. Therefore, thermal treatment seems to be a very promising plasma effect enhances the photoresponse range and absorption capac­
method to adjust the crystal phase of the photocatalyst by controlling its ity and the utilization efficiency of catalysts towards light through local
preparation conditions and improve its selectivity of the products; yet plasma resonance effect (LSPR, a phenomenon in which the free e- of
insights into the effect of crystal phases on the acidity/basicity of pho­ metal particles resonance with the incident photoelectric magnetic
tocatalyst surface requires further study. field). Secondly, the loaded metal particles, functioning as a proton
reduction site, could effectively reduce the overpotential of the photo­
catalysts. Because of the small Fermi level and large work function of
3.4. Other structural factors
noble metals, e- can be effectively transferred from the semiconductor to
the surface of metal particles and participate in proton reduction, which
In addition to the impact of band gap structure, morphology and size,
is conducive to hinder photogenerated carriers recombination rate and
and crystal phase, other structural factors influencing the biomass
improving the photocatalytic performance. Thirdly, plasma photo­
photoreforming efficiency and H2 production include: 1) Introducing
thermal effect can effectively increase the surface temperature of the
cocatalysts by depositing noble metals, metal phosphides and sulfides;

Fig. 6. Photoreforming glucose for H2 production over


3DOM TiO2-Au: (a) H2 production over different photo­
catalysts; (b) XPS Au 4f spectra of 3DOM TiO2-Au before and
after photocatalytic reaction; (c) H2 generation over 3DOM
TiO2-Au before and after treated by H2/Ar under 500 ◦ C. The
photocatalyst before treatment is labeled as 3DOM TiO2-
Au+/Au0 and the photocatalyst after treatment is labeled as
3DOM TiO2-Au0; (d) Glucose consumption, arabinose pro­
duction and arabinose selectivity over 3DOM TiO2-Au under
visible light (λ > 420 nm); (e) Schematic illustration for the
arabinose and gas co-production over 3DOM TiO2-Au.
Reprinted with permission from ref. [45], Copyright 2021
Elsevier.

8
C. Shi et al. Chemical Engineering Journal 452 (2023) 138980

photocatalyst, thus increasing the specific reaction activity [86]. oxidation of benzyl alcohols and H2 production in anaerobic water [91].
The use of plasma shows obvious effectiveness in improving the The NixSy nanoparticles showed the enhanced ternary effects in H2 yield,
biomass photoreforming efficiency, as it can facilitate depolymerization alcohol conversion, and benzaldehyde selectivity, the H2 yield over ZnS-
of cellulose and lignin by cleaving the β-1, 4 glycosidic bond and β-O-4 NixSy (3.65 mmol g-1 h− 1) was 5 times than ZnS (0.73 mmol g-1 h− 1).
bond, respectively. The produced small molecules can be further adop­ Considering the rareness of noble metals, environmental-friendly and
ted for photoreforming reaction [41]. Zhao et al prepared gold nano­ inexpensive transition metals phosphides and sulfides have great po­
particles decorated three dimensional ordered macroporous TiO2-Au tential in photoreforming field.
(3DOM TiO2-Au) for glucose photoreforming for arabinose and gas fuel
(H2, CH4, and CO) co-production [45] (Entry 9 in Table 1). The catalyst 3.4.2. Enhanced selective adsorption of photocatalysts toward substrates by
gave about 37 % glucose conversion and ended up with 0.2 g L-1 of introducing transition metals
arabinose production after 8 h reaction, which were 2.8 and 10 times Compared with the plasma effect of high cost and rare noble metals,
higher than that of using conventional TiO2, respectively. For the gas transition metals (such as Mn, Fe, Co, Ni, and Cu) play a more prominent
products, the H2 production of 3DOM TiO2-Au showed a linear increase role by regulating their selectivity to biomass substrates for the
in the first 3 h at a rate of 10.90 mmol g− 1 h− 1, then gradually plateaued improved photoreforming efficiency. Ni is one of the most extensive
after 7 h reaction (Fig. 6a). This may be due to the consumption of H2 by researched transitions metals in photoreforming. It has been identified
the CO hydrogenation reaction to generate CH4. The XPS study of 3DOM that metallic Ni is an excellent catalyst for H2 production while oxidized
TiO2-Au before and after the reaction showed that a large amount of Au Ni species can facilitate the oxidation reactions of organics [31,42,92].
(I) present before the reaction was reduced to Au(0) during the reaction Han et al. reported an ultrathin CdS nanosheets modified with co-
(Fig. 6b), which can be ascribed to the reducing effect of the H2 and CO catalyst Ni (Ni/CdS) to upgrade furfural alcohol and 5-HMF to value-
that were generated in situ. Meanwhile, the LSPR effect of Au NPs could added furfural and DFF with concomitant formation of H2 in neutral
also promote the reduction of Au(I). As a result, the reduction of Au(I) water under visible light irradiation [42] (Entry 10 in Table 1). Aided by
content may further decrease the arabinose selectivity and H2 produc­ theoretical computation, they found that the slightly stronger binding
tion, which was confirmed by a series of controlled experiments. Firstly, affinity of the aldehyde group in 5-HMF with water/NiO (0 0 1) interface
the synthesized 3DOM TiO2-Au was pretreated with H2/Ar mixture at in Ni/CdS resulted in the lower transformation of HMF to DFF compared
500 ◦ C to reduce Au(I) to Au(0), to evaluate its performance discrepancy to that of furfural alcohol to furfural in neutral water. This also caused
in the production of arabinose and H2 during glucose photoreforming. the lower H2 production by photoreforming 5-HMF than furfural
All the glucose conversion, arabinose, and H2 production (Fig. 6c) were alcohol. Furthermore, Han et al. prepared M/CdS (M = transition metals
lower after H2/Ar treatment. In addition, results showed that the glucose in the first row of the periodic table, including Mn, Fe, Co, Ni, and Cu)
conversion under visible light was negligible, while the arabinose pro­ for photoreforming lignin model compounds to small aromatic products
duction exhibited almost linear relationship with reaction time and the and H2 [31]. Firstly, the photoreforming behaviors of the M/CdS sam­
selectivity was maintained as high as 80 % (Fig. 6d). This further sup­ ples were tested by benzylic alcohol oxidation. As given in Fig. 7a,
ported that Au NPs were the key active sites for arabinose and gas compared to pristine CdS nanosheets, the overall performance of M/CdS
products production from glucose photoreforming (Fig. 6e). Similarly, samples were substantially improved, including benzylic alcohol con­
Hippargi et al. prepared Au-Pt/TiO2 for photoreforming acetic acid to version, benzaldehyde yield, and H2 evolution. Among various types of
produce H2 and methane [87]. Under UV–Visible light (400 W) irradi­ transition metal loaded catalysts, Ni/CdS achieved the most favored
ation, the newly prepared Au-Pt/TiO2 reported effective production of performance, with a 93.4 % conversion of benzylic alcohol, nearly 100
H2 and methane at the rate of 9.39 and 2.21 mmol h− 1 (14.73 and 62.58 % of benzaldehyde selectivity, 93.4 % of benzaldehyde and H2 yield.
mmol g-1 h− 1), respectively, from acetic acid with apparent quantum Interestingly, unlike the traditional photoreforming reactions where H2
yield of 3.85 % at the bandgap of 3.2 eV. Overall, the plasma effect can are sourced from water, in this system H2 was derived from benzyl
promote the reforming reaction of biomass effectively, but the rareness alcohol, which was evidenced by comparing the relationship between
of noble metals reduces their value in commercial applications. benzyl alcohol consumption and H2 production with different solvents
Metal phosphides and sulfides (like Ni2P, CoP, and NixSy etc.) can (water and acetonitrile). It was found that the benzyl alcohol con­
also work as an appropriate candidate for cocatalyst in photoreforming sumption and H2 production were consistent, regardless of type of sol­
[88–91]. Zhang et al. demonstrated an environmental-friendly CoP/ vent used. In addition, a control experiment was carried out by using
Zn2In2S5 catalyst for highly selective photoreforming methanol to parent CdS without Ni to demonstrate the significant role of Ni species.
generate ethylene glycol (EG) and H2 under visible light and AM1.5 The corresponding 1H NMR spectra showed only a small amount of
irradiation [88]. α–C–H in methanol was first activated to form hydrobenzoin was obtained while no benzaldehyde was detected, which
⋅CH2OH, then C–C coupling occurs and got EG. Furthermore, it was first clearly proved the critical role of Ni species in the photocatalytic
reported the activity of Zn2In2S5 system for photoreforming ethanol to oxidation of benzyl alcohol to benzaldehyde on Ni/CdS (Fig. 7b).
value-added 2, 3-butanediol and H2 under visible or solar light irradi­ Considering the source of H, the efficiency of H2 production by photo­
ation (the oxidation pathway of ethanol: α–C–H activation to obtain reforming benzyl alcohol was correspondingly reduced.
⋅CHCH3OH and C–C coupling). Tasleem et al. constructed a 2D Ti3AlC2
MAX dispersed TiO2 heterostructure with Ni2P as a cocatalyst for effi­ 3.4.3. Modification of acid-base sites
cient photoreforming H2 production in the presence of methanol–water Apart from the impact of catalyst crystal phase on the acidity and
and glycerol-water mixture, accompanied by the production of CO2 with alkalinity of its surface, surface modification of acidic or basic functional
a significant yield [90]. TiO2/Ni2P/Ti3AlC2 composites exhibited the groups also has an effect on it. Surface acid-base sites of photocatalyst
enhanced activity (1.29 times) than TiO2/Ti3AlC2, which could be play a crucial role in biomass photoreforming, since it likely affects the
attributed to the significant increase in charge separation by Ni2P as depolymerization of the macromolecular biomass substrates and the
cocatalyst. In addition, glycerol-water mixture showed higher potential accessibility between the photocatalysts and the biomass [93,94]. Hao
for H2, CO2, and CH4 production due to the generation of more photon et al. prepared chemisorbed SO2- 4 modified P25-NixSy composites for H2
and OH ions. Uekert and coworkers also confirmed the potential of Ni2P evolution by photoreforming cellulose (Fig. 7c) [94] (Entry 11 in
as cocatalyst in photoreforming. They fabricated an inexpensive and Table 1). The H2 production rate was found to be 3.02 mmol g-1h− 1 at
nontoxic carbon nitride/nickel phosphide (CNx/Ni2P) photocatalyst to 80 ◦ C, which was 76 times higher than that of Pt/P25. This can be
reform poly(ethylene terephthalate) (PET) and poly(lactic acid) (PLA) to attributed to the positive effect of the surface SO2-4 on the hydrolysis of
H2 and various organic chemicals under alkaline aqueous conditions cellulose into small glucose molecules, which served as e- donor and
[89]. Loading NixSy cocatalyst on ZnS nanorods enabled the selective diproton donor in photoreforming reaction. Meanwhile, protons

9
C. Shi et al. Chemical Engineering Journal 452 (2023) 138980

Fig. 7. Photoreforming benzyl alcohol for H2 generation over M/CdS: (a) Activity comparison of M/CdS photocatalysts for the oxidation of benzyl alcohol; (b) 1H
NMR spectra of benzylic alcohol oxidation before and after photocatalysis using pure CdS as the photocatalyst. Reprinted with permission from ref. [31], Copyright
2019 American Chemical Society. SO2-4 modified P25-NixSy for photoreforming H2 evolution from cellulose: (c) Diagram H2 production by cellulose photoreforming
into over SO2-
4 modified P25-NixSy under solar irradiation. Reprinted with permission from ref. [94], Copyright 2018 Wiley.

reduction was accomplished on NixSy to generate H2. In terms of H2 photocatalyst and biomass substrates. The crystal phase of the photo­
production from biomass reforming, the acid-base sites modified on the catalyst mainly affects the photoreforming H2 production efficiency by
catalyst surface mainly trigger biomass hydrolysis, isomerization of affecting its substrate selectivity and the acidity and alkalinity of the
corresponding derivatives and dehydration. Recently, Wang et al. catalyst surface. By adjusting the type and proportion of crystal phase of
showed that both proton and e- transfer can be enhanced by creating photocatalyst, the composition of various liquid and gas phase products
SO2- 2-
4 on the surface of CdS. The bifunctional SO4 serves as proton can be effectively regulated, which is an important way to control the
acceptor to promote proton transfer and can increase the oxidation selectivity and yield of biomass photoreforming reactions.
potential of the VB to enhance e- transfer. SO2-
4 /CdS can convert a variety
of saccharides (glucose, fructose, maltose, sucrose, xylose, lactose, in­ 4. Influences of reaction conditions on photoreforming H2
sulin, and starch) to syngas. For glycerol without CO2, the CO generation generation efficiency
rate (0.31 mmol g-1 h− 1) and the H2 generation (0.05 mmol g-1 h− 1) of
SO2-
4 /CdS were 9 and 4-fold higher than that of pristine CdS, respectively Apart from designing favored photocatalyst structure as discussed in
[95]. However, the role of SO2- 4 in simultaneously promoting both pro­ Section 3, properly adjusting photoreforming reaction conditions could
ton and e- transfer during photocatalytic biomass reforming remains to improve the performance of photocatalyst in photoreforming biomass
be further explored. Brønsted acid, commonly employed in biomass producing H2. In the following sub-sections, the effect of reaction pa­
photoreforming reactions, is generally used as a catalyst for acid depo­ rameters (light intensity, the concentration of photocatalyst and sub­
lymerization of macromolecular biomass. However, more attempts are strate, temperature, pH, the involved metal salt solutions, reactor
needed to investigate the role of Brønsted acid in other photoreforming configuration, and solvent type) will be elaborated with representative
reactions. Up to now, no substantive work has been reported on the examples on biomass photoreforming H2 production activity, which are
application of Lewis acid in biomass photoreforming. In addition, the summarized in Table 2.
interaction between the Lewis acid/base and the Brønsted acid/base
sites also needs to be demonstrated by relevant work. 4.1. Effect of light
In conclusion, the photocatalyst structure (band gap, morphology,
size, and crystal phase) has a great impact on its photoreforming per­ Light provides energy to move forward photochemical reactions.
formance. The band gap structure mainly affects the optical properties Factors including the type of light source, spectral range (ultraviolet
and redox ability of the catalyst, which directly affect the light utiliza­ (UV), visible, or near IR light) and irradiation intensity are all needed to
tion efficiency of the catalyst, the reaction pathway and degree of the be considered. Generally, the selection of spectral range is based on the
substrates. The morphology and size of photocatalyst may affect the light absorption characteristics of the photocatalyst, which is strongly
species of active free radicals and the accessibility between linked with the band gap of catalysts discussed in Section 3. As a type of

10
C. Shi et al. Chemical Engineering Journal 452 (2023) 138980

Table 2
Influences of reaction conditions on photoreforming H2 generation efficiency.
Influence factor Entry Photocatalyst Substrate Light source, range Temperature H2 evolution rate AQY/% Reference
(Concentration, g L-1) (Concentration, g L-1) (nm), and intensity (◦ C)/pH (mmol g-1h− 1) (λex/nm)
(mW cm− 2)

Effect of light 1 CdS/MoS2 (1.00) Glucose (18.02) Xe lamp, λ>400, (300 55.00 – [101]
W)
NCN
2 CNx (0.17) 4-MBA (1.22) Xe lamp, AM1.5 (1 0 0) 34.48 – [103]
3 Pt/TiO2 (1.33) Methanol (3 mM) Near-UV, λmax = 365 – [107]
(0.50) ~5.63*10-5
(1.00) ~8.77*10-5
Methanol (1.90) ~0.00014
(100 mM) (3.00) ~0.00017
(0.50) ~0.00017
(1.90) ~0.00034
(2.30) ~0.00044
(3.00) ~0.00102
Photocatalyst 4 Pt/TiO2 (0.17) Methanol (3 mM) Near-UV, ~4.39*10-5 – [107]
concentration (0.33) λmax = 365 (3.00) ~7.94*10-5
(0.67) ~0.00013
(1.33) ~0.00016
(2.64) ~0.00019
(4.00) ~0.00019
(5.00) Methanol ~0.00021
(0.33) (100 mM) ~0.00026
(1.33) ~0.00050
(4.00) ~0.00054
5 Ru-doped LaFeO3 Glucose UV, 375<λ<380, (57) – [109]
(0.75) (1.00) ~0.49
(1.00) ~1.05
(1.50) ~1.30
(2.00) ~1.15
(3.00) ~0.55
Substrate 6 CdS/CdOx (3.00) α-cellulose (50.00) Xe lamp, AM1.5 (1 0 0) 2.57 – [21]
concentration Hemicellulose
(25.00)Lignin 2.32
(0.25)
0.53
7 Pt-TiO2 (1.33) Fructose Xe arc lamp (300 W) T: 40 0.00017 – [112]
(0.20 mM) T: 60 0.00023
8 TiO2/NiOx (0.20) Cellulose (20.00) Xe lamp T: 25 0.27 –
Reaction (300 W) T: 80 4.00
temperature 9 Pt/TiO2 (1.00) Corn stover (30.00) Xe lamp T: 30 ~0.15 – [117]
(300 W) T: 45 ~0.22
T: 60 ~0.27
10 Platinized TiO2 (0.20) Cellulose (0.05) UV–vis (250 W) T: 50 0.13 – [41]
T: 130 1.32
pH 11 Pt-F-TiO2 (1.50) Glucose (2.00) UV-LEDs strip (10 W) pH 2 0.59 – [81]
12 o-g-C3N4 (0.25) Glucose (0.1 M) Solar light (50) pH 11 1.37 – [69]

abundant and renewable energy, sunlight contains about 5 % UV light For evaluation of photocatalyst, solar simulator is often used as a
(wavelength less than 400 nm), 40 % visible (400–800 nm), and 50 % IR convenient and stable radiation source for photoreforming reactions.
light (>800 nm). For semiconductors with broad band gap such as TiO2 Xenon lamp is a common type of solar simulators that can emit
(>3.2 eV), the photoreforming activity can only be triggered by high continuous wavelengths, achieving the standard AM 1.5 average solar
energy photon illumination, such as UV light [24,41,81,94,96–100]. As radiation at ground level in temperate zones by adding filters (air mass
discussed above, more efficient use of sunlight can be realized by 1.5 global filter). The light absorbed by the reaction system is signifi­
designing materials that can initiate photocatalytic reforming activity cantly critical for driving forward the photoreforming reaction, which is
under low-energy photon radiation (visible light), like CdS and g-C3N4 correlated with the applied light intensity [10]. A lot of research has
based materials [17,21,36,42,69,101–104]. Li et al. synthesized a binary been done on the influence of the applied light intensity on the photo­
heterostructure CdS/MoS2 flowerlike composite photocatalyst by a catalytic reaction [105–108]. Nomikos et al. studied the kinetics and
simple one-pot hydrothermal method [101] (Entry 1 in Table 2). When mechanism of photoreforming methanol reaction over Pt/TiO2 irradi­
the mass ratio of CdS and MoS2 reached 40: 100, the photocatalyst ated under near-UV light (λmax = 365 nm) [107] (Entry 3 in Table 2).
showed a superb H2 evolution rate of 55.00 mmol g-1 h− 1 with glucose as The influence of incident light intensity (I0) on the reaction rate was
the sacrificial agent under visible light irradiation. Kasap et al. reported investigated in the range of 0.5–3.0 mW cm− 2 with a constant Pt/TiO2
a cyanamide-functionalized carbon nitride, NCNCNx, which was used to photocatalyst concentration (1.33 g L-1). It was observed that an in­
produce H2 via photoreforming under visible-light irradiation in the crease of I0 from 0.5 to 2.3 mW cm− 2 resulted in an increase of the
presence of a molecular Ni bis (diphosphine) cocatalyst (NiP) in aqueous maximum H2 production rate (rmax) at lower initial methanol concen­
media (Fig. 8a) [103] (Entry 2 in Table 2). The photoreforming H2 tration in solution (3 mM). With further increase of I0 to 3.0 mW cm− 2,
production rate can achieve 34.48 mmol g-1 h− 1 in potassium phosphate rmax did not improve significantly (Fig. 8d). For solution carrying higher
solution (KPi, pH 4.5) at 25 ◦ C with 4-methyl benzylalcohol (4-MBA) as initial methanol concentration (100 mM), the rmax increases mono­
easily oxidized model cellulose substrate, 1.12 mmol g-1 h− 1 with tonically with increase of I0 (Fig. 8e). Furthermore, they studied the
glucose, and 0.20 mmol g-1 h− 1 even with sawdust, a raw and unpro­ dependence of rmax on I0 for two different initial methanol concentra­
cessed biomass samples, under the same conditions (Fig. 8b-c). tions. The results showed that the rmax was proportional to I0 for 100 mM

11
C. Shi et al. Chemical Engineering Journal 452 (2023) 138980

Fig. 8. Photoreforming of lignocellulose into


H2 using nanoengineered carbon nitride under
benign conditions: (a) Schematic representation
of H2 generation through photoreforming of
lignocellulose with NCNCNx and H2 production
cocatalysts; Photocatalytic H2 formation using
activated and bulk NCNCNx (0.5 mg) with NiP
(300 nmol) in KPi (0.1 M, pH 4.5, 3 mL) with
(b) 4-MBA (30 μmol) and (c) purified lignocel­
lulose components (100 mg) under AM1.5G
irradiation at 25 ◦ C. Reprinted with permission
from ref. [103], Copyright 2018 American
Chemical Society. Photoreforming methanol
over Pt/TiO2: Effect of incident light intensity
[I0: a. 0.5, b. 1.0, c. 1.9, d. 2.3, and e. 3.0 mW
cm− 2] on the H2 evolution rate for Pt/TiO2
(Ccat. = 1.33 g/L) of initial methanol concen­
trations of (d) 3.0 mM and (e) 100 mM; (f) The
relationship between the rmax and I0 at different
initial methanol concentrations (a. 3.0, and b.
100 mM) obtained from (d) and (e). Reprinted
with permission from ref. [107], Copyright
2014 Elsevier.

methanol concentration whereas the rmax follows a linear relationship between H2 generation rates and the amount of Pt/TiO2 photocatalyst in
with the square-root of I0 for 3 mM methanol concentration, which photoreforming methanol tests, showing an optimal photocatalyst load
indicated that the dependence of the heterogeneous photocatalytic re­ of 1.33 g L-1 [107], which supported the hypothesis (Entry 4 in Table 2).
actions rate on incident light intensity can be approximated by either In photoreforming glucose for H2 evolution using Ru-doped LaFeO3,
linear or square-root functions (Fig. 8f). Iervolino and co-workers discovered the optimal photocatalyst loading
of 1.5 g L-1 [109] (Entry 5 in Table 2). In the photoreforming system
4.2. Photocatalyst concentration using C3N4 as photocatalyst, the optimal H2 evolution rate was obtained
at 0.16 g L-1 photocatalyst concentration with α-cellulose and purified
With sufficient incident photons, the required concentration of the lignocellulose components as substrates [103]. In essence, the optimal
photocatalyst mainly depends on the absorption capacity of photo­ photocatalyst concentration, for which the whole photocatalyst surface
catalyst to incident light, and the surface active sites present for biomass is effectively irradiated, should be analyzed in combination with the
photoreforming. At lower photocatalyst concentration, the biomass reaction conditions and the geometrical characteristics of the light
photoreforming reaction is controlled by the amount of surface active source-photoreactor system [109,111].
sites of photocatalyst. Therefore, the photoreforming rate increases with
increase of photocatalyst concentration. The maximum photoreforming 4.3. Substrates concentration
rate can be achieved when the photocatalyst concentration reaches the
saturation limit, at which point the photocatalysts can utilize the largest Photocatalytic biomass or biomass-derivatives reforming for H2
proportion of photons irradiated into the catalytic system. Further production generally requires the participation of water, in which
increasing the catalyst’s concentration beyond such saturation value biomass or biomass derivatives acting as h+ scavengers consume
leads to reduced photoreforming rate. Decreased photoreforming rate is oxidizing active species to obtain a series of liquid products, and where
generally observed at high photocatalyst concentrations, possibly due e- reduces H+ to produce H2. It has been identified that properly tuning
to: 1) The reduced light absorption by light scattering and reflection of the substrate concentration can achieve the maximum substrate acces­
photocatalyst particles as the opacity of reaction system increases; and sibility, which is conducive to achieve improved photoreforming activ­
2) The decreased amount of active sites due to particles agglomeration ity and selectivity. Therefore, substrate concentration is a very
[41,97,99,103,107,109,110]. Nomikos et al. reported the relationships important factor in photocatalytic process. In the following sections, the

12
C. Shi et al. Chemical Engineering Journal 452 (2023) 138980

effects of different substrates and different substrate concentrations on In the presence of CdS/CdOx, in addition to signal i, a new signal of
the photoreforming H2 production rate and products selectivity are polymeric carboxylic acid resonance (δ = 182 ppm, ii) appeared, which
summarized. corresponded to the oxidized polysaccharide end groups. However, the
Many studies focused on the photoreforming glucose or other sharp monosaccharide peaks introduced by cellulose decomposition did
monosaccharides [96,109,112], and it is worthwhile to note that an not appear in the spectrum, which may have been oxidized rapidly and
increase of the H2 evolution was observed with up to 1000 mg L-1 of completely into end products (Fig. 9a). It was confirmed that the pho­
initial glucose concentration [113]. The analysis of the liquid products tocatalyst exhibited better performance for the oxidation of mono­
showed that gluconic acid was the only chemical in the liquid phase as saccharides by photoreforming 0.1 M glucose, and its H2 generation rate
the glucose photoreforming product, and a small amount of CO2 was was 130 % higher than that of α-cellulose.
further produced by the decarboxylation of gluconic acid. As glucose Due to the different nature of substrate (structural complexity, light
concentration further increases, both the glucose degradation and absorption, and solubility), the same study has showed that the optimal
photoreforming H2 evolution rate decreased [109]. A similar trend has concentrations for hemicellulose and lignin are 25 and 0.25 g L-1,
been observed in photoreforming alcohols [107] and polysaccharides showing H2 generation rates of 2.32 and 0.53 mmol g-1 h− 1, respectively
[15,21,41,99,100,103] for H2 production. (Fig. 9b). As a control, CdS/CdOx without cellulose (Fig. 9c, black line)
Wakerley et al. used wood-derived highly crystalline α-cellulose to present showed only 10 % of the H2 evolution activity for the first 24 h
test the photoreforming H2 production activity of CdS/CdOx in 10 M and then ceased (4 % after six days), which originated from oxidation of
KOH under simulated solar light (AM 1.5G, 100 mW cm− 2) [21] (Entry 6 residual N, N-Dimethylformamide (DMF) from the CdS stock solution
in Table 2). Due to the low solubility of α-cellulose, optimization [21]. The same conclusion can be drawn based on the study of homo­
experiment revealed that a photocatalyst dosage of 50 g L-1 led to the geneous photoreforming system for H2 generation [15].
highest photocatalytic activity (2.57 mmol g-1 h− 1). By examining the
13
C-labeled cellulose oxidation products using a 13C NMR spectroscopy,
the signals of polymeric cellulose resonances (δ = 55–110 ppm, i) were 4.4. Reaction temperature
found to show up under irradiation without photocatalyst, which
confirmed that α-cellulose can be partially dissolved under basic con­ In terms of overcoming the activation energy to move forward
dition, and the solvated cellulose acts as a substrate for photocatalysis. biomass photoreforming reactions, increasing temperature helps
improve biomass photoreforming efficiency. Many mechanisms have

Fig. 9. Solar-driven reforming of lignocellulose


to H2 with CdS/CdOx: (a) 13C NMR spectra of the
supernatant from a suspension of uniformly 13C-
labelled cellulose (10 mg) in 10 M NaOD in D2O
(1 mL) after three days of irradiation with and
without 1 nmol of CdS/CdOx QDs; (b) The rate of
H2 generation through photoreforming ligno­
cellulose (50 mg mL− 1 α-cellulose, 25 mg mL− 1
hemicellulose (xylan from beech wood), and
0.25 mg mL− 1 lignin over CdS/CdOx; (c) Long-
term photocatalytic cumulative production of
H2 by CdS/CdOx QDs (0.5 µM) with and without
50 mg mL− 1 α-cellulose in 2 mL KOH (10 M); (d)
The pH dependence of the photocatalytic H2
evolution from a 2 mL aqueous solution of
ligand-free CdS QDs (0.5 µM) with 50 mg mL− 1
α-cellulose after 18 h of irradiation. Reprinted
with permission from ref. [21]. Copyright 2017
Nature. Photocatalytic H2 generation on Ni-
decorated CdS nanorods: (e) Schematic of the
photocatalytic generation of H2 under the pro­
posed hole shuttle mechanism; (f) Energy dia­
gram for the two-step oxidation reaction
(https://doi.org/10.1038/NMAT4049). Reprin­
ted with permission from ref. [58]. Copyright
2014 Nature. Photoreforming of biomass in
metal salt hydrate solutions: (g) Photo-reforming
results after 24 h using different photocatalysts
and conditions using AM 1.5G, 100 mW cm− 2
irradiation at 25 ◦ C. Cellulose (50 mg) in 1 mL of
62.5 wt% LiBr in 0.1 M H2SO4 is added to 1.5 mL
aqueous solution containing 4 mg photocatalyst
to give 2.5 mL of 25 wt% LiBr. Reprinted with
permission from ref. [114]. Copyright 2020
Royal Society of Chemistry.

13
C. Shi et al. Chemical Engineering Journal 452 (2023) 138980

been proposed to expound the effects of reaction temperature on pho­ (which can be measured as zeta potential) of the photocatalyst surface,
toreforming H2 production: Firstly, in case that the irradiation flux is chemical state of the substrate, adsorption or chemisorption processes
intense enough to maintain a significant population of photogenerated (interaction between substrates and photocatalyst), redox potentials of
e--h+ in photocatalysts, the dependence of photoreforming H2 produc­ both substrate molecules and protons, stabilization of either reaction
tion efficiency on temperature is affected by behaviors such as the intermediates or products, or particle aggregation
diffusion/adsorption of substrates and desorption of H2 [69,81,110,112–114,120]; 2) The influence on the semiconductor band
[99,112,115,116]; Secondly, increasing temperature is conducive to gap structure [58,72]; 3) The influence of pH on the solubility of
lignocellulose dissolution, which can provide sufficient e- donors for the lignocellulose components with complex structure, as varying pH may
photocatalytic reaction and promote H2 generation affect the substrate accessibility, and then affects the photoreforming H2
[41,94,103,111,117–119]. Therefore, it is strongly important to pay production efficiency [15,21,41,100,103,117,118]. Considering the
attention to the specific reaction system when analyzing the effects of complex influence of pH on the photocatalytic system (the properties of
reaction temperature on photoreforming H2 production. photocatalysts, substrates and products, proton redox potential and
By employing biomass-derived oxygenated substrates (glucose or different photocatalyst/substrate combinations), predicting the general
fructose) as e- donors, Kondarides et al. studied the photoreforming H2 trends is a complicated task, thus we will describe the influence of re­
production performance on Pt/TiO2 photocatalyst at different reaction action pH on the photoreforming H2 generation rate.
temperatures [112] (Entry 7 in Table 2). It was clearly demonstrated
that the beneficial influence of higher temperature on reaction rate, 4.5.1. Effects on photocatalyst’s surface properties
which is likely due to the acceleration of “dark” catalytic reaction steps In Fu and co-workers’ work about photoreforming glucose using
(such as oxidation of oxygenated substrates by photogenerated O2) powdered TiO2-based photocatalyst in water, they clearly demonstrated
occurring on the Pt surface, which was not directly associated with the that the correlation between acidity/basicity of the reaction medium
light-induced reaction steps. It is also possible that increasing temper­ and the H2 production [81,110,112,113,120]. The effect of solution pH
ature also affected the mass transfer of fructose to the photocatalyst on the H2 evolution is not straightforward, involving changes of the
surface and/or the desorption of H2, thereby enhancing the apparent H2 catalyst surface charge, the chemical state of glucose (undissociated in
evolution rate in the gas-phase. It was observed that H2 production rates the pH range investigated (pKa ca. 12.3) and the oxidation reduction
could be enhanced by almost 50 % when increasing the reaction tem­ potential of H+/H2. The different H2 yields obtained could be explained
perature from 40 to 60 ◦ C (from 0.00017 to 0.00023 mmol g-1 h− 1), by a combined impact of these different factors [10]. In one of the re­
while a less noticeable effect was observed at 80 ◦ C. It should be pointed ported works, they pointed out that a weak basic condition was more
out that the temperature increase will affect the H2 production rate but favorable for photocatalytic H2 production, while acidic condition was
will not affect the overall H2 production of the system. The research of disadvantageous as shown by the low H2 generation efficiency in the pH
Chiarello [115], Lan [99], and Kondarides [116] et al. also supported range (from 1 to 5) [113]. The glucose in the solution was mainly in
that the light-driven reaction steps were not affected by a slight tem­ molecular form and would prefer to bond with undercoordinated surface
perature change. In contrast, photoreforming sub-processes that possess Ti atoms through its hydroxyl O when the pH < pKa (the pKa of glucose
small activation energy, such as adsorption/desorption dynamics, ma­ is about 12.3). Increasing solution pH (in the range of pH < pKa) will
terials diffusion, dark oxidation steps involving photogenerated charges facilitate the adsorbed glucose to become deprotonated into RCH2–O-
or stabilization of transient intermediates, are more likely to be accel­ that has been proven to capture h+ more efficiently than the neutral
erated by mild heating. molecule, thus, resulting in improved photoreforming performance.
As pointed out in Section 4.3, substrate accessibility is an important When further increasing pH to exceed the glucose’s pKa, the TiO2 sur­
prerequisite for photocatalytic biomass reforming for H2 production, face was negatively charged because the pH value is higher than the
and increasing temperature is beneficial to not only improve material isoelectric point of TiO2 (pH = 6), and the glucose is also negatively
diffusion but also promote lignocellulose dissolution into smaller, more charged to introduce electrostatic repulsion that will hinder the contact
accessible fragments. This may account for the improved photo­ between glucose and TiO2, leading to decreased H2 production rate
reforming H2 evolution of certain lignocellulose components with [113].
complex structures (Fig. 1) when the temperature was elevated Iervolino et al. reported the H2 evolution by photoreforming glucose
[41,94,111,117,119]. Zhang et al. reported an enhanced H2 generation over Pt-F-TiO2. Different from the work by Fu et al., the H2 production
by photoreforming cellulose catalyzed by a graphitic carbon layer rate in Iervolino et al.’s study was found to increase with decreasing
loaded TiO2/NiOx nanoparticles [119] (Entry 8 in Table 2). When the initial solution pH, and can reach the maxima of about 0.59 mmol g-1
temperature rose from room temperature to 80 ◦ C, the H2 production h− 1 at pH = 2 [81] (Entry 11 in Table 2). A different mechanism was
rate increased from 0.27 to 4.00 mmol g-1 h− 1. Zhou et al. studied the hypothesized to expound such discrepancy. In acidic condition, the
photocatalytic alkali pretreated corn stover photoreforming on Pt/TiO2, redox potential of H+/H2 would become more positive, and this condi­
showing enhanced H2 evolution when temperature was raised to 60 ◦ C tion was advantageous for efficient H2 generation. For different photo­
[117] (Entry 9 in Table 2). Furthermore, Zou and co-workers combined catalysts (such as g-C3N4 and CdS), the optimal pH of photoreforming H2
photocatalysis and acid hydrolysis to simultaneously achieve biomass production is also different, suggesting that it is a case-specific param­
conversion and H2 production [41] (Entry 10 in Table 2). Photo­ eter. For example, Speltini et al. reported an increased H2 production
reforming H2 generation was accomplished by cellulose hydrolysis in was observed by either decreasing or increasing pH to more acidic or
0.6 M sulfuric acid solution in the presence of a photocatalyst (such as more basic, and the highest and the lowest H2 evolution rates were
platinized TiO2) under UV-light irradiation. Results showed that H2 achieved at pH = 11 and 7, respectively [69] (Entry 12 in Table 2). This
generation efficiency at 403 K was over 10-folds higher than that at 323 is mainly due to the difference in surface charge of the photocatalyst and
K, suggesting less sufficient acid hydrolysis of cellulosic feedstock at 323 the chemical state of glucose under different pH conditions.
K. In conclusion, temperature needs to be optimized individually for a
specific reaction. 4.5.2. Effects on photocatalyst’s band gap structure positions
The band gap positions of semiconductor may change with pH,
4.5. Effect of initial pH which will affect overall photoreforming efficiency. Simon’s work has
shown that increasing pH led to a cathodic shift in the CB and VB po­
The medium pH has been identified to impact the following aspects tential of CdS. Since the VB potential of CdS is higher than that of
of the biomass photoreforming process: 1) Effects on the properties of ⋅OH/–OH, the –OH can be oxidized to ⋅OH with stronger oxidation
photocatalysts, including the surface charge and electrokinetic potential ability (when pKa = 11.9, ⋅OH is deprotonated), thus promoting the

14
C. Shi et al. Chemical Engineering Journal 452 (2023) 138980

oxidation of the substrate. This was claimed to be the reason for the MSH containing 0.1 M H2SO4 into 1.5 mL of aqueous 0.1 M LiOH instead
higher ethanol photoreforming rate at high pH (the optimal H2 gener­ of pure H2O, and the H2 production efficiency can reach 0.11 mmol g-1
ation rate is 63.00 mmol g-1 h− 1 when pH = 14.7) (Fig. 9e-f) [58]. h− 1 (Fig. 9g). All these studies demonstrated that MSH is a suitable
Meanwhile, the photocorrosion of CdS can be effectively suppressed medium for lignocellulosic biomass photoreforming, as those lignocel­
under strongly alkaline conditions. This result was different from the lulosic biomasses can be depolymerized under relatively mild condi­
previous Tsubomur’s report [121], which indicated that the flat band tions, converting them into arabinose, erythrose, and formic acid that
potential of CdS is consistent at different pH conditions, but acidic are easily accessible to colloidal photocatalysts in photoreforming
conditions is more conducive to H+ reduction as it gives stronger process.
negative polarity. For TiO2 materials [122,123], its VB position follows The selectivity of substrate is influenced by the type of solvent used
the standard Nernstian behavior, which is parallel to the oxidation and in photoreforming reaction. In general, water is the preferred solvent for
reduction potential of water. photocatalytic oxidation because it offers many practical advantages,
It should be noted that the aforementioned behaviors must be also such as the environmental friendliness and the availability of protons.
taken into account the substrate nature, particularly for those with or However, considering the improved selectivity and stability of liquid
without acidic functional groups as the effect of pH could be dramati­ products, the reduced ionization effects of the solvents on the photo­
cally different on them. For acidic substrates, such as oxalic acid [124], catalyst surface, and the prolonged lifetime of radical specious (like
formic acid [125], lactic acid [126], etc., the photoreforming H2 pro­ singlet oxygen) in solvent, organic solvents such as acetonitrile
duction rate decreases with increasing pH. This may be due to the in­ [97,127], and tetrahydrofuran [128,129] etc., are also considered as
hibition of substrate adsorption on the negatively charged suitable medium for biomass photoreforming reaction. Reactor config­
semiconductor surface. Similarly, for substrates without acidic func­ uration is also a factor that affects the photocatalytic reforming activity.
tional groups, behaviors including that increased H2 production with Iervolino at al. investigated the effect of the reactor configuration with
increasing alkalinity, as well as the widespread oxidation of –OH into the optimized photocatalyst [109]. It is observed that the percentage of
corresponding free radicals (⋅OH), have been recognized in many reports irradiated photocatalyst volume is the key factor, which is affected by
[10]. Sanwald and co-workers indicated that changing pH could affect the inner diameter of the reactor. When the reactor volume is consistent,
the photoreforming rate by influencing the hydrolysis of the reaction the percentage of irradiated photocatalyst in reactor with smaller inner
intermediate [72]. For example, formates can be converted into the diameter is higher than that of larger one. These results are in agreement
corresponding aldose intermediate through light-driven, redox-neutral with a previous work [130].
hydrolysis under pH-neutral and acidic conditions. The dynamics of this In summary, photoreforming reaction conditions affect the proper­
reaction are slow, which requires the interaction of substrates with ties of both photocatalyst and substrate, thus exerting more influence on
negative and positive photogenerated charges, resulting in blockage at the H2 production efficiency. In the actual biomass photoreforming, it is
the active sites of photoanode and severe e--h+ recombination. Under necessary to optimize these parameters collectively for maximum H2
alkaline conditions, stable H2 evolution and saccharides conversion evolution and biomass conversion.
were achieved by rapid hydrolysis and cracking of formic acid in­
termediates induced by OH–. 5. Photoreforming lignocellulose and its derivatives

4.5.3. Effect on the solubility of lignocellulose substrates Lignocellulose is the most abundant biomass in nature. It has a
Improved substrates solubility and valorization rate can be achieved complex multi-component structure, which provides it strong mechan­
by properly altering the system pH. Wakerley et al. proposed that the ical and chemical stability [131], as shown in Fig. 1. The main compo­
photoreforming H2 evolution activities were improved with increasing nent of lignocellulose is cellulose, which is composed of β-1, 4-glycosidic
higher solubility of cellulose under alkaline conditions [21]. The bond linked anhydroglucose units and stabilized by intra- and inter-
dependence of H2 evolution on pH (1.0, 2.5, 5.0, and 10.0 M KOH) was molecular hydrogen bonds to form strong and insoluble elementary fi­
studied in a 2 mL solution of ligand-free CdS QDs (0.5 µM) with 50 mg brils. Hemicellulose is a branched copolymer derived from different
mL− 1 α-cellulose substrate, demonstrating larger volumes of H2 were types of pentoses and hexoses, which serves as a crosslinker to inter­
produced at higher pH. Besides, they further coated the CdS surface with connect cellulose elementary fibrils and lignin. Lignin is an amorphous
CdOx monolayer or multilayer, which provided a thin surface- polyether compound formed by connecting phenylpropane units
passivating shell and improved the stability of CdS at high concentra­ through C–C and ether bonds. It is derived from phenolic monomers (p-
tions of KOH (Fig. 9d). A similar result was also achieved by Kasap coumaryl alcohol, coniferyl alcohol, and sinapyl alcohol) of different
[103], Achilleos [15], and Zhou et al. [117] In addition, there are also contents. Assembled by plant cells, cellulose, hemicellulose, and lignin
some reports showing that acidic conditions are conducive to lignocel­ form recalcitrant lignocellulosic matrix that is resistant to external
lulose hydrolysis [41,118], all aiming at improving the substrate physical and chemical treatment [81]. Direct photocatalytic lignocel­
accessibility and promoting H2 production rate. lulose reforming under ambient conditions is challenging, and there is a
great incentive to optimize photocatalyst structure and photocatalytic
4.6. Other reaction conditions system for enhanced lignocellulose photoreforming performance. Here,
the recent progresses in photoreforming H2 generation using different
Other reaction conditions, such as the use of metal salt hydrate so­ biomass-derived substrates, including cellulose, hemicellulose, lignin,
lutions (MSH), reactor configuration, and the type of solvent used in raw biomass, and small molecules derived from biomass (including
photoreforming H2 production have also been investigated. Pichler et al. glucose, furfurals, and organic acid, alcohols, etc.), are critically
demonstrated the use of MSH as a reaction medium for the photo­ reviewed, which are shown in Table 3.
catalytic lignocellulose reforming using different types of photocatalysts
such as TiO2 and cyanamide-functionalized carbon nitride (NCNCNx) 5.1. Cellulose and hemicellulose
[114]. TiO2 photocatalysts showed a much higher H2 production rate in
LiBr MSH (62.5 wt% LiBr in 0.1 M H2SO4) in photoreforming cellulose Photoreforming of cellulose and hemicellulose into chemicals and
compared to that conducted in 2 M H2SO4. A possible reason, as claimed clean biofuels has attracted great interests due to the abundant feedstock
by the researchers, was that LiBr MSH benefited the dissolution and availability [132]. Compared to mono- or disaccharides, high efficiency
depolymerisation of cellulose and made them more readily to partici­ photoreforming of cellulose or hemicellulose requires sufficient acces­
pate the photoreforming process. In particular, they explored basic sibility between the photocatalysts and substrates, which poses a chal­
conditions for photoreforming and added 1 mL cellulose LiBr (62.5 wt%) lenge for the combination of insoluble polymer substrates with

15
C. Shi et al. Chemical Engineering Journal 452 (2023) 138980

Table 3
Photoreforming activity of different lignocellulose substrates.
Photocatalyst Entry Substrate Substrate pretreatment Light source, intensity (mW H2 evolution AQY/% (λex/ Reference
cm− 2) rate nm)
(mmol g-1h− 1)

CdS/CdOx QDs and Co 1 α-cellulose 10 M KOH Solar light simulator (1 0 0) 2.30 – [21]
(BF4)2 Xylan 25 ◦C 2.05
Lignin 0.26
Raw biomass 5.30
Rutile TiO2 2 Cellulose 62.5 wt% LiBr in 0.1 M Solar light simulator 0.18 – [114]
H2SO4) (1 0 0)
Pt/TiO2 3 Cellulose 0.6 M H2SO4 Iron doped halide lamp (250 1.32 – [41]
W)
Polymeric CN 4 Arabinose 5 M NaOH Xe lamp (300 W) 0.71 7.87 [135]
Xylan 1.17 (4 2 0)
Xylose 4.09
Galactose 5.21
NiP/NCNCNx 5 α-cellulose 0.1 M KPi (pH 4.5) 25 ◦C Solar light simulator (1 0 0) 0.25 – [103]
xylan 0.14
Cellobiose 0.51
Glucose 1.12
Xylose 0.55
Galactose 0.76
Sinapyl alcohol 0.04
Lingin 0.02
NiS/CdS 6 Lignin – Xe lamp 1.51 44.9 [17]
Eosin Y/Pt/TiO2 7 Olive mill wastewater pH = 3 Solar box (25) 0.18 – [138]
CoO/g-C3N4 8 Regenerated cellulose 1 M NaOH Xe lamp (300 W) 0.18 – [43]
II
Pt/TiO2 9 Glucose Hg lamp 0.21 [113]
Soluble starch Microwave (125 W) 0.19
Pt/(CNT-TiO2) 10 Arabinose – Heraeus TQ 150 medium 0.09 – [46]
Fructose pressure Hg lamp 0.05
Glucose 0.10
Cellobiose 0.08
SNGODs 11 Sucrose pH = 10 Xe lamp (300 W) 0.22 11.00 [142]
Glucose 0.16 (4 2 0)
Ni/CdS 12 5-HMF – Blue LED 0.03 – [42]
CdS/MoS2 13 5-HMF – Xe arc lamp (110.30) 27.16 – [153]
Zn0.5Cd0.5S/ 14 5-HMF – White LED bulbs (30 W) 0.06 24.5(4 5 0) [155]
MnO2
Ni-Cu-TiO2 15 Glycerol – Mediumpressure Hg lamp 1.62 – [157]
(450 W)

heterogeneous photocatalysts [133,134]. As described in Section 3 and 4, cellulose photoreforming process (such as extreme alkaline or acidic
for these types of substrates, the optimization of photocatalyst structure media for complete substrate dissolution) requires intensive energy
and reaction system is particularly critical. consumption and the use of chemical-resistant, anti-corrosive reactor, it
Many studies have been conducted to improve the accessibility of may not be suitable for large-scale production. To achieve a milder
cellulose and hemicellulose substrate. 1) Optimizing the solubility of photoreforming process, Reisner et al. incorporated MSH for biomass
cellulose and hemicellulose helps enhance their accessibility to photo­ photoreforming [114], as the lignocellulosic biomass can be dissolved in
catalyst. To achieve so, it is usually necessary to tune the reaction MSH solution under relatively mild conditions (Entry 2 in Table 3). The
conditions such as the solvent system and pH (strong acid or base soluble polysaccharides can be easily exposed to colloidal photocatalysts
environment) as mentioned in Section 4. 2). Hydrolysis of cellulose and for H2 production during the photoreforming process. The cellulose
hemicellulose to produce biomass fragments with improved accessi­ solution treated with MSH was used for photoreforming reaction over
bility, which can be facilitated by the introduction of solid acids into the TiO2 nanoparticles, reporting a maximum H2 evolution rate of over 0.18
photocatalytic system [43]. mmol g-1 h− 1 (Fig. 9g).

5.1.1. Dissolved cellulose and hemicellulose 5.1.2. Hydrolyzed cellulose and hemicellulose
To improve the solubility of cellulose, it is usually required to Zou et al. employed acid hydrolysis to hydrolyze cellulose for
conduct the reaction under strong acid or alkali environment. As improved accessibility in biomass conversion and H2 production [41]
described in Section 4.5.3, Wakerley et al. tested the catalytic efficiency (Entry 3 in Table 3). The photoreforming process was carried out in 0.6
of CdS/CdOX quantum dots on photoreforming α-cellulose under high M sulfuric acid solution at 403 K under ultraviolet light irradiation, in
concentration of KOH (10 M) solution [21] (Entry 1 in Table 3). The the presence of a platinized TiO2 (Pt/TiO2), achieving an efficient and
yield of H2 evolution in such conditions was about 2.30 mmol g-1 h− 1. stable H2 production of 1.32 mmol g-1 h− 1. This also implied the stability
Isotope labeling test proved that the H2 released by the photocatalyst of photocatalyst in such strong acidic condition. Mono- or oligosaccha­
was mainly from the reduction of H2O, as the ratio of D2 to HD produced rides were produced in situ by acid hydrolysis of cellulose and then
by photocatalysis in D2O was 4:1, and the formation of HD was due to served as substrate for photoreforming. In detail, glucose was produced
the H/D exchange with the OH group of the cellulose. In the same as the main hydrolysate, while other hydrolysates such as fructose
research, the reforming of hemicellulose (xylan from beech wood) was (isomer of glucose), 5-HMF (dehydration from fructose), and other
also discussed in parallel, indicating xylan can be simultaneously pho­ carbohydrates were also detected. According to HPLC analysis of the
toreformed to give a similar H2 production yield (over 2.00 mmol g-1 liquid products, the cellulose transformation pathway during photo­
h− 1). Considering that the harsh reaction conditions of the most reforming was determined as shown in Fig. 10a. Liu et al. studied the H2

16
C. Shi et al. Chemical Engineering Journal 452 (2023) 138980

generation by efficient photoreforming of different lignocellulose over quantum efficiency (AQE) of H2 production being 44.9 %. When lignin
polymeric carbon nitride (PCN). As displayed in Fig. 10b, the photo­ and lactic acid were simultaneously used as h+ scavenger, a variety of
reforming activity of cellulose hydrolysates, cellobiose and glucose, was liquid products including methanol, ethanol, formaldehyde, formic acid,
significantly higher than that of cellulose. Photoreforming of arabinose, and oxalic acid can be generated. The h+ reaction, following the
xylan, xylose, and galactose enable the HER of 0.71, 1.17, 4.09, and pathway I and II in the Fig. 10c was the rate-limiting step in the whole
5.21 mmol g-1 h− 1, respectively. The results demonstrated the difference photocatalytic reaction. Alternatively, lignin can be photoreformed in
of H2 generation value in the presence of different lignocellulosic com­ alkaline aqueous solution using CdS quantum dots [21]. The smaller
ponents may be limited by the accessibility to trap the photoinduced h+ band gap of CdS/CdOX made it capable of absorbing visible light with
in PCN [135] (Entry 4 in Table 3). wavelengths above 420 nm. Since lignin showed a wide absorption peak
Due to high degree of cellulose polymerization and the robust inter- at 300 nm and a shoulder peak at 350 nm, there is no competition be­
and intra-molecular hydrogen bonding, the application of photo­ tween the photocatalyst and the lignin in the utilization of light sources,
catalytic degradation of cellulose and hemicellulose reforming for H2 guaranteeing improved photoreforming efficiency. At lignin loadings of
production still needs more research. A simple, energy saving, and mild 0.25 mg mL− 1, CdS/CdOX catalyst was able to evolve H2 at a yield of
reaction condition is necessary to depolymerize cellulose into smaller 0.26 mmol g-1 h− 1. Nevertheless, most of the studies on photoreforming
saccharide molecules. It is also important to change the structure of lignin is in the preliminary stage (Entry 1 in Table 3). The complex and
cellulose through certain physiochemical pretreatment to improve its diverse structural characteristics of lignin still make it difficult to un­
photoreforming efficiency. dergo photocatalytic reforming process.

5.2. Lignin 5.3. Raw biomass and biomass wastes

As another major component of lignocellulose, lignin has attracted a H2 production from photoreforming raw biomass and biomass waste
great deal of attention due to its abundance and renewable properties. can be considered more efficient biomass utilization than isolated and
Compared to cellulose and hemicellulose, the structure of lignin is more purified components, as no separation and purification process is
complex, diverse, and stable. In addition, the brown color of lignin may required. This is a long-term and challenging goal for advancing biomass
affect the absorption of light by photocatalyst. Therefore, photo­ conversion technology, with relatively few research at present.
reforming lignin to produce H2 is relatively more difficult. That being It has been demonstrated that raw biomass, such as wood, grasses
said, some aromatic hydrocarbons can be generated in the lignin pho­ stem (wheatgrass [136], straw, and fescue grass [137], etc.), and cotton
toreforming for H2 evolution process, which implies highly important [21] can be used for H2 evolution by photoreforming. Reisner et al.
industrial value. investigated the H2 generation activity by photoreforming raw biomass
Considering the complexity of the lignin structure, researchers usu­ with size less than 0.25 cm by coarse cutting using CdS/CdOX and CO
ally start with model lignin compound (sinapyl alcohol) to study its (BF4)2 co-catalysts in 2 mL KOH (10 M) (Fig. 5c) [21]. >5 mmol g-1 h− 1
photoreforming H2 production activity, though it is questionable of H2 production yield was observed using wood from branches, while
whether they perform alike. In a study focusing on activated NCNCNx and other raw biomass (sawdust and grass) showed lower activity (Fig. 11a).
NiP catalysts, photoreforming H2 production from sinapyl alcohol gave a In addition, some biomass waste, like bagasse, waste paper, cardboard,
rate of 0.04 mmol g-1 h− 1, while the yield was only 0.002 mmol g-1 h− 1 newspaper, and olive mill wastewater (OMW) have also been used for
from lignin under same condition [103] (Entry 5 in Table 3). In another direct photoreforming H2 evolution [21,138]. Up to 0.18 mmol g-1 h− 1
study, Li et al. synthesized a uniform one-dimensional NiS/CdS nano­ of H2 productivity can be achieved from photoreforming OMW by Pt/
composite for H2 evolution [17] (Entry 6 in Table 3). A maximum H2 TiO2 (Entry 7 in Table 3). Therein, ⋅OH and h+ were considered as the
production yield of 1.51 mmol g-1 h− 1 was achieved when photo­ main oxidizing species to convert biomass. Meanwhile, the polyphenols
reforming both lactic acid and lignin as h+ scavengers, with apparent fraction in OMW, showing strong antioxidant properties, can act as the

Fig. 10. Photoreforming cellulosic


biomass waste to H2 by merging pho­
tocatalysis with acid hydrolysis: (a)
Proposed pathway of cellulose decom­
position in combined processes of hy­
drolysis and photocatalysis. Reprinted
with permission from ref. [41]. Copy­
right 2018 Elsevier. H2 generation by
efficient photoreforming of different
lignocellulose over polymeric carbon
nitride (PCN): (b) Photocatalytic H2
evolution rates using PCN as the pho­
tocatalyst with different purified
lignocellulose components (200 mg).
Reprinted with permission from ref.
[135]. Copyright 2020 Elsevier. H2
evolution by photoreforming of lignin
and lactic acid over NiS/CdS nano­
structures: (c) Scheme for the photo­
catalytic H2 production over NiS/CdS
nanocomposites under visible light.
Reprinted with permission from ref.
[17]. Copyright 2018 Elsevier.

17
C. Shi et al. Chemical Engineering Journal 452 (2023) 138980

Fig. 11. Photocatalytic H2 evolution from crude other biomass derivatives over CdS/CdOx: (a) Photoreforming 50 mg mL− 1 of raw and waste biomass substrates for
H2 generation on CdS/CdOx. Reprinted with permission from ref. [21]. Copyright 2017 Nature. Photoreforming of pinewood (Pinus ponderosa) acid hydrolysate for
H2 generation: (b) GCMS and HPLC identification of compounds in the hydrolysate and AC-treated hydrolysate. Reprinted with permission from ref. [118]. Copyright
2017 Elsevier.

sacrificial agents to promote H2 release from water [138]. H2 production rate increased significantly by 81 % (i.e, 0.33 mmol g-1
As discussed earlier, photoreforming of raw biomass and biomass h− 1 in H2 productivity) [138]. Wu et.al investigated the photoreforming
waste is generally less active for H2 generation, pretreatment is also performance of raw lignocellulosic biomass wheat straw (WS), with/
essential to make better use of them. Reisner et al. employed high without various biomass pretreatment strategies by using engineered
alkaline conditions (10 M KOH) to provide an in-situ pretreatment of the CoO/g-C3N4 [43] (Entry 8 in Table 3). It was found the pretreated
wood from branches for photoreforming H2 production [21]. Jaswal biomass had much higher H2 production and cellulose conversions as
et al. transformed ground pines into hydrolysate (containing sugars, compared with raw biomass, but the degree of improvement is highly
carboxylic acids, furfuryl acids, and phenols, Fig. 11b) using diluted acid depended on pretreatment strategies.
for photoreforming H2 generation [118]. The hydrolysate was treated Overall, high-value chemicals and clean fuels can be produced
with activated carbon (AC), which removed phenols, followed by pho­ through photoreforming of raw and biomass waste, which is in line with
toreforming with 1 % Pt/TiO2. Under optimized reaction conditions, 8 the requirements of sustainable development. However, despite being a
consecutive photocatalytic cycles of the hydrolysates produced a cu­ promising substrate for H2 production by photoreforming, the poor
mulative H2 yield of 19.9 mL g− 1 pinewood. In addition, Speltini et al. catalyst accessibility makes raw biomass a less efficient substrate for
used magnesium silicate (125 g L-1, 6 h exposure) to pretreat OMW. The photoreforming H2 production. More attention should be paid on it to

18
C. Shi et al. Chemical Engineering Journal 452 (2023) 138980

realize carbon neutrality. photoconversion of glucose in aqueous solution under aerobic and
anaerobic conditions [96]. The Pt-brookite catalyst produced about
5.4. Small molecules derived from biomass 1.70 mmol of H2 after 7 h of radiation in anaerobic condition, and the
photoreaction resulted in isomerization of glucose into fructose, which
Compared with the complex structure of cellulose, hemicellulose, was further oxidized into high-value gluconic acid, arabinose, erythrose,
lignin, and raw biomass, small molecules derived from biomass and formic acid. The loaded Pt was found to dominate the rate and
(monosaccharide and disaccharides, furfurals, organic acid, and alco­ selectivity of brookite reforming glucose for H2 production. Wu et al.
hols etc.) have less stringent requirements for photocatalysts and cata­ carried out photoreforming glucose over M/TiO2 catalyst (M = Pt, Rh,
lytic systems, so they have also been studied more widely. Ru, Ir, Au, Ni, and Cu) to improve the selectivity of gas products, i.e.,
producing H2 with low CO concentration (Fig. 12b) [150]. The selec­
5.4.1. Monosaccharide and disaccharides tivity and the generation of H2 were largely dependent on the kinds of
Monosaccharide and disaccharides are the most widely studied metals deposited on TiO2. Among the tested, Rh/TiO2 catalyst gave a
model substrates for biomass photoreforming, due to their abundancy in high H2 production with the lowest CO concentration, with the product
lignocellulose (i.e., cellulose and hemicellulose). They are ideal e- do­ selectivity increasing in the order of Ir < Ru < Au < Ni≈Cu≈Pt < Rh.
nors for photocatalytic H2 production. Monosaccharides mainly include In addition to glucose, other monosaccharides (fructose and arabi­
five-carbon sugars (xylose [139,140] and arabinose [46]) and six-carbon nose, etc.) and some disaccharides, like cellobiose, have also been used
sugars (glucose [141,142], fructose [46,143,144], galactose, and in photoreforming. Silva et al. prepared carbon nanotube-TiO2-T com­
mannose [145]), as shown in Fig. 12a. Disaccharides that has been used posites ((CNTs-TiO2)-T, T annotating the temperature of calcination in
as biomass reforming substrates include cellobiose [46,146], sucrose Kelvin) for H2 production from photoreforming of arabinose, glucose,
[38,113,147,148], maltose [116,149], lactose [112], etc. Glucose is the fructose, and cellobiose [46]. For the control photocatalyst (Pt/TiO2-
most basic unit that constitutes cellulose and accounts for the most 473), the H2 production efficiency increases following the order of
significant part of the sugar monomers of lignocellulose, while the other cellobiose, fructose, glucose, and arabinose, appearing to be related to
sugars are mainly from hemicellulose. the complexity of the carbon skeleton [113,151] (Entry 9 in Table 3).
Glucose is widely studied for photocatalytic H2 production. Bellar­ For biomass substrate with more complex carbon skeleton (glucose,
dita et al. used Pt-loaded TiO2 as the photocatalyst and studied the fructose, and cellobiose), the photoreforming efficiency of Pt/(CNT-

Fig. 12. (a) Structural formulae of several mono- and di-saccharides molecules. Reprinted with permission from ref. [11]. Copyright 2018 Wiley. (b) Schematic
model of the photoreforming of glucose on metal/TiO2 catalysts yielding H2 production with low CO selectivity. (c) Conceptual illustration of a graphene oxide dot
(GOD) for photocatalytic reforming of sugar and glucose into H2. Reprinted with permission from ref. [142]. Copyright 2019 Royal Society of Chemistry.

19
C. Shi et al. Chemical Engineering Journal 452 (2023) 138980

TiO2)-473 was higher than that of Pt/TiO2-473. Especially for photo­ To photoreform 5-HMF under mild conditions, one promising way is
reforming cellobiose, the H2 generation rate of Pt/(CNT-TiO2)-473 was to design heterogeneous and the use of non-metallic photocatalysts.
2.8 folds of that for Pt/TiO2-473 after 2 h of light irradiation (Entry 10 in Battula et al. synthesized a novel porous carbon nitride (SGCN) system,
Table 3). This may be related to the higher availability of e- at Pt par­ which could accomplish the H2 evolution form H2O and the DFF pro­
ticles and the higher surface area of the composite (61 m2 g− 1), which duction from selective HMF oxidation [154]. 12 μmol h− 1 m− 2 of H2 was
resulted in the improved H+ reduction and reduced diffusion limit in the achieved after 6 h visible light irradiation while the DFF yield of 13.8 %
case of larger molecules, respectively. Nguyen et al. reported an envi­ with > 99 % selectivity. As runs longer (48 h), the selectivity of DFF was
ronmentally friendly catalyst consisting of sulfur and nitrogen co-doped maintained > 99 %, an improved DFF yield 38.4 %, and 36 μmol h− 1
GO dots (SNGODs) to effectively reform sugar (primarily sucrose) and m− 2 of H2 production rate. Moreover, 6.2 μmol h− 1 m− 2 of H2 with 7.2 %
glucose (Fig. 12c) [142] (Entry 11 in Table 3). These functionalized DFF yield were produced under natural sunlight for 6 h. Dhingra et al.
SNGODs expanded the e- resonance domain, narrowed the band gap, synthesized Zn0.5Cd0.5S/xMnO2 (x represents the weight ratio of MnO2
and led to charge delocalization and separation. The SNGODs exhibited loaded on Zn0.5Cd0.5S) heterojunction for photoreforming 5-HMF for
higher photocatalytic activity of 0.22 and 0.16 mmol g− 1 h− 1 of H2 DEF and H2 production (Entry 14 in Table 3). Among all the formula­
production from sucrose and glucose, respectively. The basal plane of tions analyzed, Zn0.5Cd0.5S/1%MnO2 showed the optimal H2 produc­
the SNGODs has a high concentration of quaternary N that repairs the tivity of 0.06 mmol g-1 h− 1 [155]. DFF being the oxidation product of 5-
vacancy defects. The functional groups effectively adsorb deprotonated HMF was supported by GC–MS results, giving a yield of 46 % after 24 h.
glucose molecules through hydrogen-bonding in a basic environment. In The time-dependent 1H NMR and HPLC analysis further confirmed the
another study, Kasap investigated the photoreforming H2 evolution oxidation pathway from 5-HMF to DFF.
using NCNCNx upon ultrasonication [103]. Within 24 h of irradiation, H2
were produced from glucose at 1.12 mmol g− 1 h− 1 with activated 5.4.3. Other biomass derivatives
NCN
CNx and NiP (Entry 5 in Table 3). When used to photoreform xylose, Other biomass derivatives that have been investigated for photo­
galactose monosaccharides, and polymeric xylan, the H2 generation reforming H2 production mainly include organic acids (carboxylic acid
rates were recorded as 0.17, 0.25, and 0.04 mmol g− 1 h− 1 within 24 h of and lactic acid, etc.), alcohols (monoalcohols such as methanol and
irradiation, respectively. Recently, Hu et al. group reported the cleavage ethanol; polyols such as glycerol, etc.), and aldehydes.
mechanism of cellobiose β-1, 4-glycosidic linkage [152]. They demon­ Organic acids are commonly used for photoreforming H2 evolution.
strated a all-solid-state 3DOM TiO2-Au-CdS Z-scheme heterojunction Cao et al used CdS nanorods (as a photosensitizer) modified Ni2P NPs for
materials for enhanced cellobiose photorefinery. The oxygen (from the photoreforming H2 generation with lactic acid as e- donor under visible
oxygen-containing free radicals produced during photocatalytic process light irradiation [156]. A highest H2 evolution rate was achieved at pH
of 3DOM TiO2-Au-CdS) insertion at C1 position of cellobiose followed by 3.0, which can be attributed to greater affinity between organic acid and
the elimination reaction, which oxidatively cleaves the β-1, 4-glycosidic catalyst due to restrained deprotonation of organic acid. The H2 evolu­
bond and results in gluconic acid and glucose generation. In presence of tion activity can sustain for>20 h and then slow down slightly. When
oxygen, glucose is further oxidized into gluconic acid which is subse­ additional aliquots of CdS or a mixture of CdS and lactic acid was added
quently oxidized or decarboxylated into glucaric acid or arabinose. This to the solution, the H2 production efficiency restored promptly to the
work may pave a novel way to rationally design photocatalyst to reveal initial state. The reduced H2 evolution efficiency can be attributed to the
mechanistic understanding of biomass photoreforming towards high- photocorrosion of CdS and the consumption of lactic acid.
value fuels and chemical feedstocks. Photoreforming processes for H2 Biomass derived alcohols and aldehydes are also competent sub­
production with monosaccharides and disaccharides as substrates have strates for photoreforming H2 production. Taylor et al. firstly prepared a
been extensively studied. However, the reaction mechanism and trans­ novel Ni-Cu-TiO2 photocatalyst for photoreforming glycerol [157],
formation pathway of photocatalytic oxidation of saccharides to high reporting a H2 productivity of 1.62 mmol g-1 h− 1, which was 44 times
value-added small molecules are still not clear or not profound. There­ than that of P25 (Entry 15 in Table 3). The Cu and Ni was found to
fore, more attention should be paid on the reaction mechanism of high significantly enhance the H2 evolution performance upon photo­
value-added products in liquid phase system while maximizing the yield reforming glycerol under near UV or visible light irradiation. Methanol
of photocatalytic H2 evolution. and ethanol can be photocatalyzed by TiO2 for H2 production [158,159].
The reaction mechanism suggests that the alcohols consume h+ and are
5.4.2. Biomass-derived furfurals transformed to organic intermediates, and the H+ get reduced to pro­
In addition to mono-, di-, and polysaccharides, another important duce H2. Interestingly, this also represents a useful approach to convert
biomass derivative used for photoreforming is furfurals. Furfurals can be aromatic alcohols into their corresponding aldehydes via photo­
incorporated in photocatalytic H2 evolution system, by serving as re­ reforming [160]. All these works provide a feasible process to utilize
action substrate for effective H2 evolution. Among all types of biomass biomass derivatives for H2 production and other value-added chemicals.
derived furfurals, 5-HMF is considered to be the most important one due It is well known that the complex aldehydes are classified as sac­
to its environmentally friendly feature and the potential of producing charides, simple aldehydes are contained at significant levels in pyrol­
value-added biomass oxidation products. The pH of the system plays a ysis bio-oils. Actually, they may make up 20-30 wt% of the total mass of
key role in the photoreforming reaction, and alkaline system is often bio-oils [10]. Utilizing aldehydes for photoreforming H2 generation is
preferred for such catalytic reaction. Simultaneous H2 production can be relatively less reported but worthy of study for the photocatalytic H2
achieved upon visible light irradiation under ambient conditions using production regarding a possible valorization approach of bio-oils. The
the biomass intermediates as h+ scavengers, which demonstrated the reduction of formaldehyde (HCHO) to H2 using H2O as one of the donors
feasibility of producing H2 from photoreforming biomass-derived in­ can be considered as an efficient and promising H2 generation strategy
termediates (5-HMF and furfural alcohol) in aqueous media [42] (Entry due to its high atomic economy and high gas purity (only H2 without CO
12 in Table 3). The H2 evolution was found to increase during the first 8 or CO2 production). Zhang et al. reported an amino functionalized
h and reach a plateau of about 4.5 mL after 8 h. Under alkaline condi­ organosilica nanotubes internally modified with AuPd nanoclusters for
tions, both furfural alcohol and 5-HMF were quickly transformed to the photoreforming H2 evolution from formaldehyde aqueous solution
corresponding acids at nearly 100 % yield. In addition, Zhao proposed [161]. Under visible light irradiation, aldehyde group was oxidized to
CdS/MoS2 nano-octahedron heterostructures for reforming furfural carboxyl group, and the H+ from H2O was reduced to H2. Photo­
alcohol under alkaline conditions, which was claimed as a cheap and reforming other aldehydes such as acetaldehyde, propyl aldehyde and
effective way [153] (Entry 13 in Table 3). The significantly enhanced benzaldehyde was investigated for H2 production as well. Besides
photocatalytic H2 evolution rates of 5.29 mmol h− 1 g− 1 were recorded. formaldehyde, other aldehydes, including acetaldehyde, propanal, and

20
C. Shi et al. Chemical Engineering Journal 452 (2023) 138980

benzaldehyde are also investigated to produce H2 [162]. biomass. The plasma effect of metal particles can promote biomass
depolymerization by increasing the catalyst’s surface temperature.
6. Conclusion and outlook Screening the pH, temperature, and other variables of the reaction sys­
tem can also be employed to determine the optimal photoreforming
In the face of environmental problems caused by global warming and condition. An ideal photocatalyst should be able to sufficiently photo­
massive use of fossil fuels, photocatalytic reforming of biomass to pro­ reform biomass in a high efficiency and mild way.
duce high energy density H2 fuel has become one of the research hot­ 2) More efforts are required in understanding the relationship be­
spots in recent years because of the environmentally friendly nature. tween HER and biomass oxidation. Generally, biomass participates in
Photoreforming biomass uses the renewable and low-cost biomass the oxidation reaction as a sacrificial agent by consuming oxidizing
sources from nature as substrate and solar light as energy source. It active species. Increasing biomass dosage may promote the availability
provides a green and sustainable way to utilize solar energy, agricul­ of active species with reductive properties, leading to improved pro­
tural, and forestry residues, suggesting its great potential in promoting duction of H2 and value-added platform compounds. The oxidation ca­
the resources circulation and mitigating environmental and energy pacity of photocatalyst will affect the oxidation reaction pathway of
pressure [163,164]. In this review, the influence of photocatalyst biomass, which in turn will influence the overall photoreforming H2
structures and photocatalytic system on H2 production from photo­ production efficiency. Therefore, the relationship between H2 genera­
reforming biomass were discussed comprehensively with representative tion and biomass oxidation should be analyzed with regard to relevant
examples. characterization methods, reaction pathways, and mechanisms.
To comprehend key structural and compositional parameters for 3) Developing photocatalyst specifically for H2 production. To date
photocatalyst, Section 3 (photocatalyst structure) focuses on discussing there are few types of photocatalysts used for photoreforming H2 pro­
the effects of band gap structure, morphology and size, crystal phases duction, and most of the studies only focus on TiO2, CdS, and bimetallic
and other factors of photocatalyst on the generation of H2 by photo­ sulfides (ZnxCd1-xS). Researchers should pay more attention to explore
reforming. Our analysis shows that all these effects are interconnected to more types of photocatalysts suitable for this field based on the reaction
some extent but can differ substantially from one system to another. In pathways and reaction mechanisms, which will play a remarkable role
addition, the interaction between H2 production and biomass oxidation in promoting the performance of biomass photoreforming to produce
by photoreforming should be analyzed in a specific reaction system; in H2.
other words, the one applied to one system may not be applicable to 4) Optimization of photocatalyst structure and photoreforming sys­
another. At the same time, much attention should be paid to the impact tem. In photoreforming biomass for H2 generation reactions, it is more
of the produced liquid products and by-products during the photo­ necessary to design photocatalyst structure based on the reaction char­
reforming process on the activity of photocatalyst. Particularly, for some acteristics in line with the reaction target. The modification of the ma­
photocatalysts like CdS, though in many studies it gave satisfactory terial’s band gap structure, crystal structure, and surface chemistry
photoreforming activity, its environmental toxicity remains to be should be conducted in consideration with its impact on the biomass
resolved. More attention should be paid to the development of catalysts oxidation half reaction, with the target of maximizing the overall pho­
with not only high photocatalytic efficiency but also nontoxic and toreforming efficiency. Many previous reports focused merely on one
environmentally compatible. For the part of photocatalytic system half reaction and may not fully reflect the whole results. Meanwhile, the
(Section 4), the influence of the initial pH, reaction temperature, and properties of the photocatalytic system also play an important but
solvent medium on the H2 production by biomass photoreforming was underrated role, which could affect the reaction path and H2 production
summarized. The influence of pH on photoreforming H2 evolution can rate as well.
be mainly summarized as three aspects: firstly, the properties of pho­ At present, there is still great discrepancy between the current art
tocatalysts, including the surface charge and electrokinetic potential and industrial expectation in terms of H2 production, liquid product
(which can be measured as zeta potential) of the photocatalyst surface, selectivity and yield. In view of the above knowledge gaps in H2 gen­
chemical state of the substrate, adsorption or chemisorption processes eration by biomass photoreforming, technology transfer from different
(interaction between substrates and photocatalyst), redox potentials of research fields is highly recommended. Novel photocatalyst structure
both substrate molecules and protons, stabilization of either reaction designed for specific reactions and capable of photoreforming raw
intermediates or products, or particle aggregation; Secondly, the influ­ biomass is of great interest to both academia and industry. Further at­
ence on the semiconductor band gap structure positions; Thirdly, the tempts are needed on investigate photocatalysts’ performance in pho­
solubility of lignocellulose components with complex structure and the toreforming biomass, as well as the reaction mechanism.
substrate accessibility. Increasing temperature usually lowers the acti­
vation energy of the rate-limiting step of photoreforming, and may also Declaration of Competing Interest
affect the type and yield of liquid products. Therefore, it is a key
parameter impacting the depolymerization of biomass macromolecules The authors declare that they have no known competing financial
into small molecules and hence directly influence the subsequent H2 interests or personal relationships that could have appeared to influence
production. Based on the current research on H2 generation form the work reported in this paper.
biomass photoreforming, we put forward the following views and
opinions: Data availability
1) Developing an appropriate depolymerization method for raw
biomass or large biomass molecules is crucial for improved photo­ No data was used for the research described in the article.
reforming H2 evolution. At present, most raw biomass undergoes a two-
step approach, i.e., depolymerization into smaller biomass derivatives, Acknowledgements
photoreforming of the biomass derivatives. However, the low conver­
sion efficiency, high cost, and complicated processing procedure limit its This work was financially supported by the National Natural Science
commercialization. For certain photoreforming systems, the production Foundation of China (No. 32071713), the Fundamental Research Funds
of by-products and contaminants make such process even more unfa­ for the Central Universities (No.2572020DX01), and the Natural Science
vorable. One potential approach to resolve these matters is properly Funds for Distinguished Young Scholars of Heilongjiang Province (No.
designing the catalyst structure and adjusting the photocatalytic system. JQ2019C001).
For photocatalyst, proper structural modification by solid acid and metal
with plasma effect may promote its performance in depolymerizing

21
C. Shi et al. Chemical Engineering Journal 452 (2023) 138980

References designs for selective oxidation reactions, Energy Environ. Sci. 14 (2021)
1140–1175.
[27] M. Estahbanati, M. Feilizadeh, F. Attar, M. Iliuta, Current developments and
[1] M. Shannon, P. Bohn, M. Elimelech, J. Georgiadis, B. Marinas, A. Mayes, Science
future trends in photocatalytic glycerol valorization: Photocatalyst development,
and technology for water purification in the coming decades, Nature 452 (2008)
Ind. Eng. Chem. Res. 59 (2020) 22330–22352.
301–310.
[28] J. Kou, C. Lu, J. Wang, Y. Chen, Z. Xu, R. Varma, Selectivity enhancement in
[2] S. Wang, B. Zhu, M. Liu, L. Zhang, J. Yu, M. Zhou, Direct Z-scheme ZnO/CdS
heterogeneous photocatalytic transformations, Chem. Rev. 117 (2017)
hierarchical photocatalyst for enhanced photocatalytic H2 production activity,
1445–1514.
Appl. Catal. B Environ. 243 (2019) 19–26.
[29] M. Qi, M. Conte, M. Anpo, Z. Tang, Y. Xu, Cooperative coupling of oxidative
[3] S. Sadhu, P. Gupta, P. Poddar, Physical mechanism behind enhanced
organic synthesis and hydrogen production over semiconductor-based
photoelectrochemical and photocatalytic properties of superhydrophilic
photocatalysts, Chem. Rev. 121 (2021) 13051–13085.
assemblies of 3D-TiO2 microspheres with arrays of oriented, single-crystalline
[30] U. Nwosu, A. Wang, B. Palma, H. Zhao, M. Khan, M. Kibria, J. Hu, Selective
TiO2 nanowires as building blocks deposited on fluorine-doped tin oxide, ACS
biomass photoreforming for valuable chemicals and fuels: A critical review,
Appl. Mater. Interfaces 9 (2017) 11202–11211.
Renew. Sust. Energy Rev. 148 (2021), 111266.
[4] S. Chen, T. Takata, K. Domen, Particulate photocatalysts for overall water
[31] G. Han, T. Yan, W. Zhang, Y. Zhang, D. Lee, Z. Cao, Y. Sun, Highly selective
splitting, Nat. Rev. Mater. 2 (2017) 17050.
photocatalytic valorization of lignin model compounds using ultrathin metal/
[5] Z. Mei, B. Zhang, J. Zheng, S. Yuan, Z. Zhuo, X. Meng, Z. Chen, K. Amine,
CdS, ACS Catal. 9 (2019) 11341–11349.
W. Yang, L. Wang, W. Wang, S. Wang, Q. Gong, J. Li, F. Liu, F. Pan, Tuning Cu
[32] H. Bahruji, M. Bowker, P. Davies, F. Pedrono, New insights into the mechanism of
dopant of Zn0.5Cd0.5S nanocrystals enables high-performance photocatalytic H2
photocatalytic reforming on Pd/TiO2, Appl. Catal. B Environ. 107 (2011)
evolution from water splitting under visible-light irradiation, Nano Energy 26
205–209.
(2016) 405–416.
[33] Z. Pan, R. Wang, J. Li, S. Iqbal, W. Liu, K. Zhou, Fe2P nanoparticles as highly
[6] N. Zhang, M. Yang, S. Liu, Y. Sun, Y. Xu, Waltzing with the versatile platform of
efficient freestanding co-catalyst for photocatalytic hydrogen evolution, Int. J.
graphene to synthesize composite photocatalysts, Chem. Rev. 115 (2015)
Hydrog. Energy 43 (2018) 5337–5345.
10307–10377.
[34] H. Zhao, C. Li, X. Yong, P. Kumar, B. Palma, Z. Hu, G. Tendeloo, S. Siahrostami,
[7] Q. Xiang, B. Cheng, J. Yu, Graphene-based photocatalysts for solar-fuel
S. Larter, D. Zheng, S. Wang, Z. Chen, M. Kibria, J. Hu, Coproduction of hydrogen
generation, Angew. Chem. Int. Ed. Engl. 54 (2015) 11350–11366.
and lactic acid from glucose photocatalysis on band-engineered Zn1-xCdxS
[8] C. Wu, S. Xue, Z. Qin, M. Nazari, G. Yang, S. Yue, T. Tong, H. Ghasemi,
homojunction, iScience 24 (2021), 102109.
F. Hernandez, S. Xue, D. Zhang, H. Wang, Z.M. Wang, S. Pu, J. Bao, Making g-
[35] I. Shkrob, T. Marin, S. Chemerisov, M. Sevilla, Mechanistic aspects of
C3N4 ultra-thin nanosheets active for photocatalytic overall water splitting, Appl.
photooxidation of polyhydroxylated molecules on metal oxides, J. Phys. Chem. C
Catal. B Environ. 282 (2021), 119557.
Nanomater Interfaces 115 (2011) 4642–4648.
[9] W. Zhen, X. Ning, B. Yang, Y. Wu, Z. Li, G. Lu, The enhancement of CdS
[36] C. Xu, P. Anusuyadevi, C. Aymonier, R. Luque, S. Marre, Nanostructured
photocatalytic activity for water splitting via anti-photocorrosion by coating Ni2P
materials for photocatalysis, Chem. Soc. Rev. 48 (2019) 3868–3902.
shell and removing nascent formed oxygen with artificial gill, Appl. Catal. B
[37] J. Liu, N. Ma, W. Wu, Q. He, Recent progress on photocatalytic heterostructures
Environ. 221 (2018) 243–257.
with full solar spectral responses, Chem. Eng. J. 393 (2020), 124719.
[10] A. Puga, Photocatalytic production of hydrogen from biomass-derived feedstocks,
[38] Q. Wang, K. Domen, Particulate photocatalysts for light-driven water splitting:
Coord. Chem. Rev. 315 (2016) 1–66.
Mechanisms, challenges, and design strategies, Chem. Rev. 120 (2019) 919–985.
[11] M. Kuehnel, E. Reisner, Solar hydrogen generation from lignocellulose, Angew.
[39] K. Christoforidis, P. Fornasiero, Photocatalytic hydrogen production: A rift into
Chem. Int. Ed. Engl. 57 (2018) 3290–3296.
the future energy supply, ChemCatChem 9 (2017) 1523–1544.
[12] M. Yin, W. Zhang, F. Qiao, J. Sun, Y. Fan, Z. Li, Hydrothermal synthesis of MoS2-
[40] N. Fajrina, M. Tahir, A critical review in strategies to improve photocatalytic
NiS/CdS with enhanced photocatalytic hydrogen production activity and
water splitting towards hydrogen production, Int. J. Hydrogen Energ. 44 (2019)
stability, J. Solid State Chem. 270 (2019) 531–538.
540–577.
[13] M. Imizcoz, A. Puga, Assessment of photocatalytic hydrogen production from
[41] J. Zou, G. Zhang, X. Xu, One-pot photoreforming of cellulosic biomass waste to
biomass or wastewaters depending on the metal co-catalyst and its deposition
hydrogen by merging photocatalysis with acid hydrolysis, Appl. Catal. A: Gen.
method on TiO2, Catalysts 9 (2019) 584.
563 (2018) 73–79.
[14] T. Kawai, T. Sakata, Conversion of carbohydrate into hydrogen fuel by a
[42] G. Han, Y. Jin, R. Burgess, N. Dickenson, X. Cao, Y. Sun, Visible-light-driven
photocatalytic process, Nature 286 (1980) 474–476.
valorization of biomass intermediates integrated with H2 production catalyzed by
[15] D. Achilleos, W. Yang, H. Kasap, A. Savateev, Y. Markushyna, J. Durrant,
ultrathin Ni/CdS nanosheets, J. Am. Chem. Soc. 139 (2017) 15584–15587.
E. Reisner, Solar reforming of biomass with homogeneous carbon dots, Angew.
[43] X. Wu, H. Zhao, M. Khan, P. Maity, T. Al-Attas, S. Larter, Q. Yong, O. Mohammed,
Chem. Int. Ed. Engl. 59 (2020) 18184–18188.
M. Kibria, J. Hu, Sunlight-driven biomass photorefinery for coproduction of
[16] N. Luo, T. Montini, J. Zhang, P. Fornasiero, E. Fonda, T. Hou, W. Nie, J. Lu, J. Liu,
sustainable hydrogen and value-added biochemicals, ACS Sustain. Chem. Eng. 8
M. Heggen, L. Lin, C. Ma, M. Wang, F. Fan, S. Jin, F. Wang, Visible-light-driven
(2020) 15772–15781.
coproduction of diesel precursors and hydrogen from lignocellulose-derived
[44] K. Sun, J. Shen, Y. Yang, H. Tang, C. Lei, Highly efficient photocatalytic hydrogen
methylfurans, Nat. Energy 4 (2019) 575–584.
evolution from 0D/2D heterojunction of FeP nanoparticles/CdS nanosheets, Appl.
[17] C. Li, H. Wang, S. Naghadeh, J. Zhang, P. Fang, Visible light driven hydrogen
Surf. Sci. 505 (2020), 144042.
evolution by photocatalytic reforming of lignin and lactic acid using one-
[45] H. Zhao, P. Liu, X. Wu, A. Wang, D. Zheng, S. Wang, Z. Chen, S. Larter, Y. Li,
dimensional NiS/CdS nanostructures, Appl. Catal. B Environ. 227 (2018)
B. Su, M. Kibria, J. Hu, Plasmon enhanced glucose photoreforming for arabinose
229–239.
and gas fuel co-production over 3DOM TiO2-Au, Appl. Catal. B Environ. 291
[18] M. Himmel, S. Ding, D. Johnson, W. Adney, M. Nimlos, J. Brady, T. Foust,
(2021), 120055.
Biomass recalcitrance: Engineering plants and enzymes for biofuels production,
[46] C. Silva, M. Sampaio, R. Marques, L. Ferreira, P. Tavares, A. Silva, J. Faria,
Science 315 (2007) 804–807.
Photocatalytic production of hydrogen from methanol and saccharides using
[19] L. Granone, F. Sieland, N. Zheng, R. Dillert, D. Bahnemann, Photocatalytic
carbon nanotube-TiO2 catalysts, Appl. Catal. B Environ. 178 (2015) 82–90.
conversion of biomass into valuable products: A meaningful approach? Green
[47] M. Huang, X. Luo, Z. Ai, Y. Li, K. Zhang, Y. Shao, B. Huang, Y. Wu, X. Hao, Band
Chem. 20 (2018) 1169–1192.
structure-controlled Zn1− xCdxS solid solution for photocatalytic hydrogen
[20] J. Shi, Y. Zou, L. Cheng, D. Ma, D. Sun, S. Mao, L. Sun, C. He, Z. Wang, In-situ
production improvement via appropriately enhancing oxidation capacity, Sol.
phosphating to synthesize Ni2P decorated NiO/g-C3N4 p-n junction for enhanced
RRL 5 (2021) 2000685.
photocatalytic hydrogen production, Chem. Eng. J. 378 (2019), 122161.
[48] A. Kurenkova, D. Markovskaya, E. Gerasimov, I. Prosvirin, S. Cherepanova,
[21] D. Wakerley, M. Kuehnel, K.L. Orchard, K.H. Ly, T.E. Rosser, E. Reisner, Solar-
E. Kozlova, New insights into the mechanism of photocatalytic hydrogen
driven reforming of lignocellulose to H2 with a CdS/CdOx photocatalyst, Nat.
evolution from aqueous solutions of saccharides over CdS-based photocatalysts
Energy 2 (2017) 17021.
under visible light, Int. J. Hydrog. Energy 45 (2020) 30165–30177.
[22] F. Parrino, M. Bellardita, E. García-López, G. Marcì, V. Loddo, L. Palmisano,
[49] Y. Zou, J. Shi, D. Ma, Z. Fan, L. Lu, C. Niu, In situ synthesis of C-doped TiO2@g-
Heterogeneous photocatalysis for selective formation of high-value-added
C3N4 core-shell hollow nanospheres with enhanced visible-light photocatalytic
molecules: Some chemical and engineering aspects, ACS Cata. 8 (2018)
activity for H2 evolution, Chem. Eng. J. 322 (2017) 435–444.
11191–11225.
[50] D. Ma, J. Shi, L. Sun, Y. Sun, S. Mao, Z. Pu, C. He, Y. Zhang, D. He, H. Wang,
[23] H. Huang, Z. Fang, K. Yu, J. Lü, R. Cao, Visible-light-driven photocatalytic H2
Y. Cheng, Knack behind the high performance CdS/ZnS-NiS nanocomposites:
evolution over CdZnS nanocrystal solid solutions: Interplay of twin structures,
Optimizing synergistic effect between cocatalyst and heterostructure for boosting
sulfur vacancies and sacrificial agents, J. Mater. Chem. A 8 (2020) 3882–3891.
hydrogen evolution, Chem. Eng. J. 431 (2022), 133446.
[24] G. Zhang, C. Ni, X. Huang, A. Welgamage, L. Lawton, P. Robertson, J. Irvine,
[51] G. Sun, J. Shi, S. Mao, D. Ma, C. He, H. Wang, Y. Cheng, Dodecylamine
Simultaneous cellulose conversion and hydrogen production assisted by cellulose
coordinated tri-arm CdS nanorod wrapped in intermittent ZnS shell for greatly
decomposition under UV-light photocatalysis, Chem. Comm. 52 (2016)
improved photocatalytic H2 evolution, Chem. Eng. J. 429 (2022), 132382.
1673–1676.
[52] S. Mao, J. Shi, G. Sun, Y. Zhang, X. Ji, Y. Lv, B. Wang, Y. Xu, Y. Cheng, Cu (II)
[25] S. Lakhera, A. Rajan, R.T.P.N. Bernaurdshaw, A review on particulate
decorated thiol-functionalized MOF as an efficient transfer medium of charge
photocatalytic hydrogen production system: Progress made in achieving high
carriers promoting photocatalytic hydrogen evolution, Chem. Eng. J. 404 (2021),
energy conversion efficiency and key challenges ahead, Renew. Sust. Energy Rev.
126533.
152 (2021), 111694.
[53] Y. Zou, J. Shi, L. Sun, D. Ma, S. Mao, Y. Lv, Y. Cheng, Energy-band-controlled
[26] C. Toe, C. Tsounis, J. Zhang, H. Masood, D. Gunawan, J. Scott, R. Amal,
ZnxCd1− xIn2S4 solid solution coupled with g-C3N4 nanosheets as 2D/2D
Advancing photoreforming of organics: Highlights on photocatalyst and system

22
C. Shi et al. Chemical Engineering Journal 452 (2023) 138980

heterostructure toward efficient photocatalytic H2 evolution, Chem. Eng. J. 378 nanocrystals: Drawing useful relations to improve the H2 yield in methanol
(2019), 122192. photosteam reforming, J. Phys. Chem. C 119 (2015) 12385–12393.
[54] H. Zhao, C. Li, L. Liu, B. Palma, Z. Hu, S. Renneckar, S. Larter, Y. Li, M. Kibria, [79] S. Song, J. Qu, P. Han, M. Hulsey, G. Zhang, Y. Wang, S. Wang, D. Chen, J. Lu,
J. Hu, B. Su, n-p heterojunction of TiO2-NiO core-shell structure for efficient N. Yan, Visible-light-driven amino acids production from biomass-based
hydrogen generation and lignin photoreforming, J. Colloid. Interface Sci. 585 feedstocks over ultrathin CdS nanosheets, Nat. Commun. 11 (2020) 4899.
(2021) 694–704. [80] J. Luo, S. Zhang, M. Sun, L. Yang, S. Luo, J.C. Crittenden, A critical review on
[55] Q. Gu, J. Long, L. Fan, L. Chen, L. Zhao, H. Lin, X. Wang, Single-site Sn-grafted energy conversion and environmental remediation of photocatalysts with
Ru/TiO2 photocatalysts for biomass reforming: synergistic effect of dual co- remodeling crystal lattice, surface, and interface, ACS Nano 13 (2019)
catalysts and molecular mechanism, J. Catal. 303 (2013) 141–155. 9811–9840.
[56] J. Nasir, Z. Rehman, S. Shah, A. Khan, I. Butler, C. Catlow, Recent developments [81] G. Iervolino, V. Vaiano, J.J. Murcia, L. Rizzo, G. Ventre, G. Pepe, P. Campiglia, M.
and perspectives in CdS-based photocatalysts for water splitting, J. Mater. Chem. C. Hidalgo, J.A. Navío, D. Sannino, Photocatalytic hydrogen production from
A 8 (2020) 20752–20780. degradation of glucose over fluorinated and platinized TiO2 catalysts, J. Catal.
[57] M. Huang, J. Yu, C. Deng, Y. Huang, M. Fan, B. Li, Z. Tong, F. Zhang, L. Dong, 3D 339 (2016) 47–56.
nanospherical CdxZn1− xS/reduced graphene oxide composites with superior [82] Q. Xu, Y. Ma, J. Zhang, X. Wang, Z. Feng, C. Li, Enhancing hydrogen production
photocatalytic activity and photocorrosion resistance, Appl. Surf. Sci. 365 (2016) activity and suppressing CO formation from photocatalytic biomass reforming on
227–239. Pt/TiO2 by optimizing anatase-rutile phase structure, J. Catal. 278 (2011)
[58] T. Simon, N. Bouchonville, M. Berr, A. Vaneski, A. Adrovic, D. Volbers, 329–335.
R. Wyrwich, M. Doblinger, A. Susha, A. Rogach, F. Jackel, J. Stolarczyk, [83] K. Li, Y. Zhang, Y. Lin, K. Wang, F. Liu, Versatile functional porous cobalt-nickel
J. Feldmann, Redox shuttle mechanism enhances photocatalytic H2 generation on phosphide-carbon cocatalyst derived from a metal-organic framework for
Ni-decorated CdS nanorods, Nat. Mater. 13 (2014) 1013–1018. boosting the photocatalytic activity of graphitic carbon nitride, ACS Appl. Mater.
[59] D. Ma, Z. Wang, J. Shi, M. Zhu, H. Yu, Y. Zou, Y. Lv, G. Sun, S. Mao, Y. Cheng, Cu- Interfaces 11 (2019) 28918–28927.
In2S3 nanorod induced the growth of Cu&In co-doped multi-arm CdS hetero- [84] C. Zhao, H. Tang, W. Liu, C. Han, X. Yang, Q. Liu, J. Xu, Constructing 0D FeP
phase junction to promote photocatalytic H2 evolution, Chem. Eng. J. 399 (2020), nanodots/2D g-C3N4 nanosheets heterojunction for highly improved
125785. photocatalytic hydrogen evolution, ChemCatChem 11 (2019) 6310–6315.
[60] J. Fang, Y. Chen, W. Wang, L. Fang, C. Lu, C. Zhu, J. Kou, Y. Ni, Z. Xu, Highly [85] L. Ding, M. Li, Y. Zhao, H. Zhang, J. Shang, J. Zhong, H. Sheng, C. Chen, J. Zhao,
efficient photocatalytic hydrogen generation of g-C3N4-CdS sheets based on The vital role of surface brönsted acid/base sites for the photocatalytic formation
plasmon-enhanced triplet-triplet annihilation upconversion, Appl. Catal. B of free ⋅OH radicals, Appl. Catal. B Environ. 266 (2020), 118634.
Environ. 258 (2019), 117762. [86] H. Nagakawa, M. Nagata, Photoreforming of lignocellulosic biomass into
[61] Y. Li, Q. Zhao, Y. Zhang, Y. Li, L. Fan, F. Li, X. Li, In-situ construction of sequential hydrogen under sunlight in the presence of thermally radiative CdS/SiC
heterostructured CoS/CdS/CuS for building “electron-welcome zone” to enhance composite photocatalyst, ACS Appl. Energy Mater. 4 (2020) 1059–1062.
solar-to-hydrogen conversion, Appl. Catal. B Environ. 300 (2021), 120763. [87] G. Hippargi, S. Anjankar, R. Krupadam, S. Rayalu, Simultaneous wastewater
[62] J. Shi, D. Sun, Y. Zou, D. Ma, C. He, X. Ji, C. Niu, Trap-level-tunable Se doped CdS treatment and generation of blended fuel methane and hydrogen using Au-Pt/
quantum dots with excellent hydrogen evolution performance without co- TiO2 photo-reforming catalytic material, Fuel 291 (2021), 120113.
catalyst, Chem. Eng. J. 364 (2019) 11–19. [88] H. Zhang, S. Xie, J. Hu, X. Wu, Q. Zhang, J. Cheng, Y. Wang, C-H activations of
[63] H. Ye, R. Shi, X. Yang, W. Fu, Y. Chen, P-doped ZnxCd1− xS solid solutions as methanol and ethanol and C-C couplings into diols by zinc-indium-sulfide under
photocatalysts for hydrogen evolution from water splitting coupled with visible light, Chem. Commun. 56 (2020) 1776–1779.
photocatalytic oxidation of 5-hydroxymethylfurfural, Appl. Catal. B Environ. 233 [89] T. Uekert, H. Kasap, E. Reisner, Photoreforming of nonrecyclable plastic waste
(2018) 70–79. over a carbon nitride/nickel phosphide catalyst, J. Am. Chem. Soc. 141 (2019)
[64] Y. Bai, Y. Zhou, J. Zhang, X. Chen, Y. Zhang, J. Liu, J. Wang, F. Wang, C. Chen, 15201–15210.
C. Li, R. Li, C. Li, Homophase junction for promoting spatial charge separation in [90] S. Tasleem, M. Tahir, Z. Zakaria, Fabricating structured 2D Ti3AlC2 MAX
photocatalytic water splitting, ACS Catal. 9 (2019) 3242–3252. dispersed TiO2 heterostructure with Ni2P as a cocatalyst for efficient
[65] H. Du, K. Liang, C. Yuan, H. Guo, X. Zhou, Y. Jiang, A. Xu, Bare Cd1-xZnxS ZB/WZ photocatalytic H2 production, J. Alloy. Comp. 842 (2020), 155752.
heterophase nanojunctions for visible light photocatalytic hydrogen production [91] H. Hao, L. Zhang, W. Wang, S. Qiao, X. Liu, Photocatalytic hydrogen evolution
with high efficiency, ACS Appl. Mater. Interfaces 8 (2016) 24550–24558. coupled with efficient selective benzaldehyde production from benzyl alcohol
[66] Y. Li, B. Sun, H. Lin, Q. Ruan, Y. Geng, J. Liu, H. Wang, Y. Yang, L. Wang, K. Tam, aqueous solution over ZnS-NixSy composites, ACS Sustain. Chem. Eng. 7 (2019)
Efficient visible-light induced H2 evolution from T-CdxZn1-xS/defective MoS2 10501–10508.
nano-hybrid with both bulk twinning homojunctions and interfacial [92] B. You, X. Liu, X. Liu, Y. Sun, Efficient H2 evolution coupled with oxidative
heterostructures, Appl. Catal. B Environ. 267 (2020), 118702. refining of alcohols via a hierarchically porous nickel bifunctional electrocatalyst,
[67] C. Prasad, H. Tang, Q. Liu, I. Bahadur, S. Karlapudi, Y. Jiang, A latest overview on ACS Cata. 7 (2017) 4564–4570.
photocatalytic application of g-C3N4 based nanostructured materials for hydrogen [93] M. Wang, J. Ma, H. Liu, N. Luo, Z. Zhao, F. Wang, Sustainable productions of
production, Int. J. Hydrog. Energy 45 (2020) 337–379. organic acids and their derivatives from biomass via selective oxidative cleavage
[68] J. Chen, Y. Ge, Y. Guo, J. Chen, Selective hydrogenation of biomass-derived 5- of C-C bond, ACS Cata. 8 (2018) 2129–2165.
hydroxymethylfurfural using palladium catalyst supported on mesoporous [94] H. Hao, L. Zhang, W. Wang, S. Zeng, Facile modification of titania with nickel
graphitic carbon nitride, J. Energy Chem. 27 (2018) 283–289. sulfide and sulfate species for the photoreformation of cellulose into hydrogen,
[69] A. Speltini, A. Scalabrini, F. Maraschi, M. Sturini, A. Pisanu, L. Malavasi, ChemSusChem 11 (2018) 2810–2817.
A. Profumo, Improved photocatalytic H2 production assisted by aqueous glucose [95] Z. Zhang, M. Wang, H. Zhou, F. Wang, Surface sulfate ion on CdS catalyst
biomass by oxidized g-C3N4, Int. J. Hydrog. Energy 43 (2018) 14925–14933. enhances syngas generation from biopolyols, J. Am. Chem. Soc. 143 (2021)
[70] X. Xu, J. Zhang, S. Wang, Z. Yao, H. Wu, L. Shi, Y. Yin, S. Wang, H. Sun, 6533–6541.
Photocatalytic reforming of biomass for hydrogen production over ZnS [96] M. Bellardita, E. García-López, G. Marcì, L. Palmisano, Photocatalytic formation
nanoparticles modified carbon nitride nanosheets, J. Colloid. Interface Sci. 555 of H2 and value-added chemicals in aqueous glucose (Pt)-TiO2 suspension, Int. J.
(2019) 22–30. Hydrog. Energy 41 (2016) 5934–5947.
[71] Q. Liu, L. Wei, Q. Xi, Y. Lei, F. Wang, Edge functionalization of terminal amino [97] L. Vià, C. Recchi, E. Yañez, T. Davies, J. Sanchez, Visible light selective
group in carbon nitride by in-situ C-N coupling for photoreforming of biomass photocatalytic conversion of glucose by TiO2, Appl. Catal. B Environ. 202 (2017)
into H2, Chem. Eng. J. 383 (2020), 123792. 281–288.
[72] K. Sanwald, T. Berto, W. Eisenreich, A. Jentys, O. Gutiérrez, J. Lercher, [98] M. Bellardita, E. López, G. Marcì, G. Nasillo, L. Palmisano, Photocatalytic solar
Overcoming the rate-limiting reaction during photoreforming of sugar aldoses for light H2 production by aqueous glucose reforming, Eur. J. Inorg. Chem. 2018
H2 generation, ACS Cata. 7 (2017) 3236–3244. (2018) 4522–4532.
[73] N. Pavlopoulos, Shining light on the role of shape-controlled nanomaterials in [99] L. Lan, Y. Shao, Y. Jiao, R. Zhang, C. Hardacre, X. Fan, Systematic study of H2
photocatalysis, Curr. Opin. Electrochem. 26 (2021), 100676. production from catalytic photoreforming of cellulose over Pt catalysts supported
[74] X. Wu, X. Fan, S. Xie, J. Lin, J. Cheng, Q. Zhang, L. Chen, Y. Wang, Solar energy- on TiO2, Chin. J. Chem. Eng. 28 (2020) 2084–2091.
driven lignin-first approach to full utilization of lignocellulosic biomass under [100] A. Speltini, M. Sturini, D. Dondi, E. Annovazzi, F. Maraschi, V. Caratto,
mild conditions, Nat. Cata. 1 (2018) 772–780. A. Profumo, A. Buttafava, sunlight-promoted photocatalytic hydrogen gas
[75] F. Pellegrino, F. Sordello, L. Mino, C. Minero, V. Hodoroaba, G. Martra, evolution from water-suspended cellulose: a systematic study, Photochem.
V. Maurino, Formic acid photoreforming for hydrogen production on shape- Photobiol. 13 (2014) 1410–1419.
controlled anatase TiO2 nanoparticles: assessment of the role of fluorides, 101}/ [101] C. Li, H. Wang, J. Ming, M. Liu, P. Fang, Hydrogen generation by photocatalytic
{001 surfaces ratio, and platinization, ACS Cata. 9 (2019) 6692–6697. reforming of glucose with heterostructured CdS/MoS2 composites under visible
[76] D. Li, R. You, M. Yang, Y. Liu, K. Qian, S. Chen, T. Cao, Z. Zhang, J. Tian, light irradiation, Int. J. Hydrog. Energy 42 (2017) 16968–16978.
W. Huang, Morphology-dependent evolutions of sizes, structures, and catalytic [102] H. Zhao, X. Ding, B. Zhang, Y. Li, C. Wang, Enhanced photocatalytic hydrogen
activity of Au nanoparticles on anatase TiO2 nanocrystals, J. Phys. Chem. C 123 evolution along with byproducts suppressing over Z-scheme CdxZn1− xS/Au/g-
(2019) 10367–10376. C3N4 photocatalysts under visible light, Sci. Bull. 62 (2017) 602–609.
[77] Y. Liu, Z. Ye, D. Li, M. Wang, Y. Zhang, W. Huang, Tuning CuOx-TiO2 interaction [103] H. Kasap, D. Achilleos, A. Huang, E. Reisner, Photoreforming of lignocellulose
and photocatalytic hydrogen production of CuOx/TiO2 photocatalysts via TiO2 into H2 using nanoengineered carbon nitride under benign conditions, J. Am.
morphology engineering, Appl. Surf. Sci. 473 (2019) 500–510. Chem. Soc. 140 (2018) 11604–11607.
[78] M. Arienzo, M. Dozzi, M. Redaelli, B. Credico, F. Morazzoni, R. Scotti, S. Polizzi, [104] J. Ma, Y. Li, D. Jin, Z. Ali, G. Jiao, J. Zhang, S. Wang, R. Sun, Functional B@mCN-
Crystal surfaces and fate of photogenerated defects in shape-controlled anatase assisted photocatalytic oxidation of biomass-derived pentoses and hexoses to
lactic acid, Green Chem. 22 (2020) 6384–6392.

23
C. Shi et al. Chemical Engineering Journal 452 (2023) 138980

[105] E. Baniasadi, I. Dincer, G. Naterer, Measured effects of light intensity and catalyst [133] B. Medronho, A. Romano, M.G. Miguel, L. Stigsson, B. Lindman, Rationalizing
concentration on photocatalytic hydrogen and oxygen production with zinc cellulose (in)solubility: reviewing basic physicochemical aspects and role of
sulfide suspensions, Int. J. Hydrog. Energ. 38 (2013) 9158–9168. hydrophobic interactions, Cellulose 19 (2012) 581–587.
[106] S. Bell, G. Will, J. Bell, Light intensity effects on photocatalytic water splitting [134] A. Gross, J. Chu, On the molecular origins of biomass recalcitrance: The
with a titania catalyst, Int. J. Hydrog. Energ. 38 (2013) 6938–6947. interaction network and solvation structures of cellulose microfibrils, J. Phys.
[107] G. Nomikos, P. Panagiotopoulou, D. Kondarides, X. Verykios, Kinetic and Chem. B 114 (2010) 13333–13341.
mechanistic study of the photocatalytic reforming of methanol over Pt/TiO2 [135] Q. Liu, F. Wang, Y. Jiang, W. Chen, R. Zou, J. Ma, L. Zhong, X. Peng, Efficient
catalyst, Appl. Catal. B Environ. 146 (2014) 249–257. photoreforming of lignocellulose into H2 and photocatalytic CO2 reduction via in-
[108] F. Concha, D. Guzmán, M. Isaacs, C. Navarrete, Operational conditions affecting plane surface dyadic heterostructure of porous polymeric carbon nitride, Carbon
formaldehyde and formic acid formation as by-products of hydrogen production 170 (2020) 199–212.
via photo-reforming of methanol using nanoparticles of TiO2, Energy Technol. 6 [136] C. Cao, L. Xu, Y. He, L. Guo, H. Jin, Z. Huo, High-efficiency gasification of wheat
(2018) 1871–1884. straw black liquor in supercritical water at high temperatures for hydrogen
[109] G. Iervolino, V. Vaiano, D. Sannino, L. Rizzo, V. Palma, Enhanced photocatalytic production, Energ Fuel. 31 (2017) 3970–3978.
hydrogen production from glucose aqueous matrices on Ru-doped LaFeO3, Appl. [137] A. Caravaca, W. Jones, C. Hardacre, M. Bowker, H2 production by the
Catal. B Environ. 207 (2017) 182–194. photocatalytic reforming of cellulose and raw biomass using Ni, Pd, Pt and Au on
[110] C. Ma, Y. Li, H. Zhang, Y. Chen, C. Lu, J. Wang, Photocatalytic hydrogen titania, Proc. Math. Phys. Eng. Sci. 472 (2016) 20160054.
evolution with simultaneous photocatalytic reforming of biomass by Er3+: YAlO3/ [138] A. Speltini, M. Sturini, F. Maraschi, D. Dondi, G. Fisogni, E. Annovazzi,
Pt-TiO2 membranes under visible light driving, Chem. Eng. J. 273 (2015) A. Profumo, A. Buttafava, Evaluation of UV-A and solar light photocatalytic
277–285. hydrogen gas evolution from olive mill wastewater, Int. J. Hydrog. Energy 40
[111] G. Ramis, E. Bahadori, I. Rossetti, Design of efficient photocatalytic processes for (2015) 4303–4310.
the production of hydrogen from biomass derived substrates, Int. J. Hydrog. [139] M. Yasuda, R. Kurogi, H. Tsumagari, T. Shiragami, T. Matsumoto, New approach
Energy. 45 (2020) 10292–10303. to fuelization of herbaceous lignocelluloses through simultaneous saccharification
[112] D. Kondarides, A. Patsoura, X. Verykios, Anaerobic photocatalytic oxidation of and fermentation followed by photocatalytic reforming, Energies 7 (2014)
carbohydrates in aqueous Pt/TiO2 suspensions with simultaneous production of 4087–4097.
hydrogen, J. Adv. Oxid. Technol. 13 (2010) 116–123. [140] T. Shiragami, T. Tomo, H. Tsumagari, R. Yuki, T. Yamashita, M. Yasuda, Pentose
[113] X. Fu, J. Long, X. Wang, D. Leung, Z. Ding, L. Wu, Z. Zhang, Z. Li, X. Fu, acting as a sacrificial multielectron source in photocatalytic hydrogen evolution
Photocatalytic reforming of biomass: A systematic study of hydrogen evolution from water by Pt-doped TiO2, Chem. Lett. 41 (2012) 29–31.
from glucose solution, Int. J. Hydrog. Energy. 33 (2008) 6484–6491. [141] Y. Li, J. Wang, S. Peng, G. Lu, S. Li, Photocatalytic hydrogen generation in the
[114] C. Pichler, T. Uekert, E. Reisner, Photoreforming of biomass in metal salt hydrate presence of glucose over ZnS-coated ZnIn2S4 under visible light irradiation, Int. J.
solutions, Chem. Commun. 56 (2020) 5743–5746. Hydrog. Energy 35 (2010) 7116–7126.
[115] G. Chiarello, D. Ferri, E. Selli, Effect of the CH3OH/H2O ratio on the mechanism [142] V. Nguyen, N. Ke, L. Nam, B. Nguyen, Y. Xiao, Y. Lee, H. Teng, Photocatalytic
of the gas-phase photocatalytic reforming of methanol on noble metal-modified reforming of sugar and glucose into H2 over functionalized graphene dots,
TiO2, J. Catal. 280 (2011) 168–177. J. Mater. Chem. A 7 (2019) 8384–8393.
[116] D. Kondarides, V. Daskalaki, A. Patsoura, X. Verykios, Hydrogen production by [143] S. Deguchi, N. Shibata, T. Takeichi, Y. Furukawa, N. Isu, Photocatalytic hydrogen
photo-induced reforming of biomass components and derivatives at ambient production from aqueous solution of various oxidizing sacrifice agents, J. Jpn.
conditions, Catal. Letters 122 (2007) 26–32. Pet. Inst. 53 (2010) 95–100.
[117] Y. Zhou, X. Ye, D. Lin, Enhance photocatalytic hydrogen evolution by using [144] R. Su, R. Tiruvalam, A. Logsdail, Q. He, C. Downing, M. Jensen, N. Dimitratos,
alkaline pretreated corn stover as a sacrificial agent, Int. J. Energy Res. 44 (2020) L. Kesavan, P. Wells, R. Bechstein, H. Jensen, S. Wendt, C. Catlow, C. Kiely,
4616–4628. G. Hutchings, F. Besenbacher, Designer titania-supported Au-Pd nanoparticles for
[118] R. Jaswal, R. Shende, W. Nan, A. Shende, Photocatalytic reforming of pinewood efficient photocatalytic hydrogen production, ACS Nano 8 (2014) 3490–3497.
(Pinus ponderosa) acid hydrolysate for hydrogen generation, Int. J. Hydrog. [145] M. Yasuda, R. Kurogi, T. Matsumoto, Quantum yields for sacrificial hydrogen
Energy 42 (2017) 2839–2848. generation from saccharides over a Pt-loaded TiO2 photocatalyst, Res. Chem.
[119] L. Zhang, W. Wang, S. Zeng, Y. Su, H. Hao, Enhanced H2 evolution from Intermed. 42 (2016) 3919–3928.
photocatalytic cellulose conversion based on graphitic carbon layers on TiO2/ [146] J. Ma, K. Liu, X. Yang, D. Jin, Y. Li, G. Jiao, J. Zhou, R. Sun, Recent advances in
NiOx, Green Chem. 20 (2018) 3008–3013. photoreforming of biomass derived feedstocks into hydrogen, biofuels, or
[120] M. Zhou, Y. Li, S. Peng, G. Lu, S. Li, Effect of epimerization of d-glucose on chemicals by using functional carbon nitride photocatalysts, ChemSusChem 14
photocatalytic hydrogen generation over Pt/TiO2, Catal. Commun. 18 (2012) (2021) 4903–4922.
21–25. [147] Y. Li, Y. Xie, S. Peng, G. Lu, S. Li, Photocatalytic hydrogen generation using
[121] M. Matsumura, M. Hiramoto, T. Iehara, H. Tsubomura, Photocatalytic and glucose and sucrose as electron donors over Pt/TiO2, Chem. J. Chin. 28 (2007)
photoelectrochemical reactions of aqueous solutions of formic acid, 156–158.
formaldehyde, and methanol on platinized CdS powder and at a CdS electrode, [148] H. Bahruji, M. Bowker, P. Davies, L. Mazroai, A. Dickinson, J. Greaves, D. James,
J. Phys. Chem. 88 (1984) 248–250. L. Millard, F. Pedrono, Sustainable H2 gas production by photocatalysis,
[122] Y. Li, G. Lu, S. Li, Photocatalytic production of hydrogen in single component and J. Photech. Photobio. A 216 (2010) 115–118.
mixture systems of electron donors and monitoring adsorption of donors by in situ [149] G. Iervolino, V. Vaiano, D. Sannino, L. Rizzo, P. Ciambelli, Production of
infrared spectroscopy, Chemosphere 52 (2003) 843–850. hydrogen from glucose by LaFeO3 based photocatalytic process during water
[123] A. Kudo, Y. Miseki, Heterogeneous photocatalyst materials for water splitting, treatment, Int. J. Hydrog. Energy 41 (2016) 959–966.
Chem. Soc. Rev. 38 (2009) 253–278. [150] G. Wu, T. Chen, G. Zhou, X. Zong, C. Li, H2 production with low CO selectivity
[124] E. Liu, L. Qi, J. Chen, J. Fan, X. Hu, In situ fabrication of a 2D Ni2P/red from photocatalytic reforming of glucose on metal/TiO2 catalysts, Sci. China, Ser.
phosphorus heterojunction for efficient photocatalytic H2 evolution, Mater. Res. B 51 (2008) 97–100.
Bull. 115 (2019) 27–36. [151] M. Cargnello, A. Gasparotto, V. Gombac, T. Montini, D. Barreca, P. Fornasiero,
[125] S. Yu, Z. Xie, M. Ran, F. Wu, Y. Zhong, M. Dan, Y. Zhou, Zinc ions modified InP Photocatalytic H2 and added-value by-products-the role of metal oxide systems in
quantum dots for enhanced photocatalytic hydrogen evolution from hydrogen their synthesis from oxygenates, Eur. J. Inorg. Chem. 2011 (2011) 4309–4323.
sulfide, J. Colloid Interface Sci. 573 (2020) 71–77. [152] H. Zhao, C. Li, X. Yu, N. Zhong, Z. Hu, Y. Li, S. Larter, M. Kibria, J. Hu,
[126] S. Cao, C. Wang, X. Lv, Y. Chen, W. Fu, A highly efficient photocatalytic H2 Mechanistic understanding of cellulose β-1, 4-glycosidic cleavage via
evolution system using colloidal CdS nanorods and nickel nanoparticles in water photocatalysis, Appl. Catal. B Environ. 302 (2022), 120872.
under visible light irradiation, Appl. Catal. B Environ. 162 (2015) 381–391. [153] L. Zhao, T. Dong, J. Du, H. Liu, H. Yuan, Y. Wang, J. Jia, H. Liu, W. Zhou,
[127] J. Colmenares, A. Magdziarz, Room temperature versatile conversion of biomass- Synthesis of CdS/MoS2 nanooctahedrons heterostructure with a tight interface for
derived compounds by means of supported TiO2 photocatalysts, J. Mol. Catal A- enhanced photocatalytic H2 evolution and biomass upgrading, Sol. RRL 5 (2021)
Chem. 366 (2013) 156–162. 2000415.
[128] X. Li, K. Peng, Q. Xia, X. Liu, Y. Wang, Efficient conversion of cellulose into 5- [154] V. Battula, A. Jaryal, K. Kailasam, Visible light-driven simultaneous H2
hydroxymethylfurfural over niobia/carbon composites, Chem. Eng. J. 332 (2018) production by water splitting coupled with selective oxidation of HMF to DFF
528–536. catalyzed by porous carbon nitride, J. Mater. Chem. A 7 (2019) 5643–5649.
[129] Z. Cao, Z. Fan, Y. Chen, M. Li, T. Shen, C. Zhu, H. Ying, Efficient preparation of 5- [155] S. Dhingra, T. Chhabra, V. Krishnan, C.M. Nagaraja, Visible-light-driven selective
hydroxymethylfurfural from cellulose in a biphasic system over hafnyl oxidation of biomass-derived HMF to DFF coupled with H2 generation by noble
phosphates, Appl. Catal. B Environ. 244 (2019) 170–177. metal-free Zn0.5Cd0.5S/MnO2 heterostructures, ACS Appl. Energy Mater. 3 (2020)
[130] V. Vaiano, O. Sacco, M. Stoller, A. Chianese, P. Ciambelli, D. Sannino, Influence of 7138–7148.
the photoreactor configuration and of different light sources in the photocatalytic [156] S. Cao, Y. Chen, C.J. Wang, P. He, W.F. Fu, Highly efficient photocatalytic
treatment of highly polluted wastewater, Int. J. Chem. React. Eng. 12 (2014) hydrogen evolution by nickel phosphide nanoparticles from aqueous solution,
63–75. Chem. Commun. 50 (2014) 10427–10429.
[131] F. Isikgor, C. Becer, Lignocellulosic biomass: a sustainable platform for the [157] S. Taylor, M. Mehta, A. Samokhvalov, Production of hydrogen by glycerol
production of bio-based chemicals and polymers, Polym. Chem. 6 (2015) photoreforming using binary nitrogen-metal-promoted N-M-TiO2 photocatalysts,
4497–4559. Chemphyschem 15 (2014) 942–949.
[132] Y. Huang, Y. Fu, Hydrolysis of cellulose to glucose by solid acid catalysts, Green [158] R. Chong, J. Li, Y. Ma, B. Zhang, H. Han, C. Li, Selective conversion of aqueous
Chem. 15 (2013) 1095–1111. glucose to value-added sugar aldose on TiO2-based photocatalysts, J. Catal. 314
(2014) 101–108.

24
C. Shi et al. Chemical Engineering Journal 452 (2023) 138980

[159] N. Lakshmana Reddy, K. Cheralathan, V. Kumari, B. Neppolian, organosilica nanotubes for visible-light-driven hydrogen evolution from
S. Venkatakrishnan, Photocatalytic reforming of biomass derived crude glycerol formaldehyde, Appl. Catal. B Environ. 220 (2018) 303–313.
in water: a sustainable approach for improved hydrogen generation using Ni [162] S. Zhang, M. Li, J. Zhao, H. Wang, X. Zhu, J. Han, X. Liu, Plasmonic AuPd-based
(OH)2 decorated TiO2 nanotubes under solar light irradiation, ACS Sustain. Chem. Mott-Schottky photocatalyst for synergistically enhanced hydrogen evolution
Eng. 6 (2018) 3754–3764. from formic acid and aldehyde, Appl. Catal. B Environ. 252 (2019) 24–32.
[160] T. Zhu, X. Ye, Q. Zhang, Z. Hui, X. Wang, S. Chen, Efficient utilization of [163] B. Xia, Y. Zhang, B. Shi, J. Ran, K. Davey, S. Qiao, Photocatalysts for hydrogen
photogenerated electrons and holes for photocatalytic redox reactions using evolution coupled with production of value-added chemicals, Small Methods 4
visible light-driven Au/ZnIn2S4 hybrid, J. Hazard. Mater. 367 (2019) 277–285. (2020) 2000063.
[161] S. Zhang, H. Wang, L. Tang, M. Li, J. Tian, Y. Cui, J. Han, X. Zhu, X. Liu, Sub 1 nm [164] X. Wu, N. Luo, S. Xie, H. Zhang, Q. Zhang, F. Wang, Y. Wang, Photocatalytic
aggregation-free AuPd nanocatalysts confined inside amino-functionalized transformations of lignocellulosic biomass into chemicals, Chem. Soc. Rev. 49
(2020) 6198–6223.

25

You might also like