Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Chinese Journal of Chemical Engineering 28 (2020) 2084–2091

Contents lists available at ScienceDirect

Chinese Journal of Chemical Engineering

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

Article

Systematic study of H2 production from catalytic photoreforming of


cellulose over Pt catalysts supported on TiO2
Lan Lan 1, Yan Shao 1,2, Yilai Jiao 3, Rongxin Zhang 1, Christopher Hardacre 1,⁎, Xiaolei Fan 1,⁎
1
Department of Chemical Engineering and Analytical Science, School of Engineering, The University of Manchester, M13 9PL, United Kingdom
2
School of Chemical and Environmental Engineering, Wuyi University, Jiangmen 52920, China
3
Shenyang National Laboratory for Materials Science, Institute of Metal Research, Chinese Academy of Sciences, Shenyang 110016, China

a r t i c l e i n f o a b s t r a c t

Article history: Hydrogen (H2) production from photocatalytic reforming of cellulose is a promising way for sustainable H2 to be
Received 1 February 2020 generated. Herein, we report a systematic study of the photocatalytic reforming of cellulose over Pt/m-TiO2 (i.e.
Received in revised form 19 March 2020 mixed TiO2, 80% of anatase and 20% of rutile) catalysts in water. The optimum operation condition was
Accepted 30 March 2020
established by studying the effect of Pt loading, catalyst concentration, cellulose concentration and reaction tem-
Available online 1 May 2020
perature on the gas production rate of H2 (rH 2) and CO2 (rCO 2), suggesting an optimum operation condition at 40
Keywords:
°C with 1.0 g·L−1 of cellulose and 0.75 g·L−1 of 0.16-Pt/m-TiO2 catalyst (with 0.16 wt% Pt loadting) to achieve a
Hydrogen (H2) production relatively sound photocatalytic performance with rH2 = 9.95 μmol·h−1. It is also shown that although the
Catalytic photoreforming photoreforming of cellulose was operated at a relatively mild condition (i.e. with an UV-A lamp irradiation at
Cellulose 40 °C in the aqueous system), a low loading of Pt at ~0.16 wt% on m-TiO2 could promote the H2 production effec-
Kinetics tively. Additionally, by comparing the reaction order expressed from both rH 2 (a1) and rCO 2 (a2) with respect to
Titanium dioxide (TiO2) cellulose and water, the possible mechanism of H2 production was proposed.
© 2020 The Chemical Industry and Engineering Society of China, and Chemical Industry Press Co., Ltd.
All rights reserved.

1. Introduction Lignocellulose is mainly comprised of three polymers which are


cellulose (N40% in wood stems), hemicellulose (20%–40%) and lignin
Lignocellulosic biomass has been regarded as the most abundant (b35%) [16]. Cellulose, the major component of lignocellulose, is a
biomass on Earth, with great potential to be converted into valuable polysaccharide of anhydroglucose monomers. As it is difficult to hy-
chemicals, such as glucose [1,2], sorbitol [3,4], lactic acid [5] and various drolyse cellulose, the valorisation of cellulose is regarded as rather
fuels such as methane [6,7], syngas (mixture of carbon monoxide, CO limited [17]. In 1980, Kawai and Sakata first discovered that plants-
and hydrogen, H2) [8,9] and H2 [6,9]. Hydrogen is an effective energy derived substrates such as sugar and cellulose could produce H 2
carrier with a high heat value of ~ 140 kJ·g−1 in comparison with under the irradiation generated by a Xe lamp. In the presence of
conventional fossil fuels such as crude oil (~ 44 kJ·g−1) and natural water and photocatalysts such as RuO2 /TiO 2 /Pt [18] and Pt/TiO2
gas (~ 39 kJ·g−1) [10], as well as being a clean energy without green- [19], H2 was generated with relatively low production rates at b 50
house gas emission from H2 combustion. It is a promising alternative μmol·h−1 ·g−1cat (as shown in Table S1). Accordingly, efforts have
to conventional fuels for the purpose of relieving the stress resulted been made to improve the H 2 production via photocatalytic
from the increasing concern in energy demand and environmental reforming of cellulose, and the relevant experimental conditions
impact [11,12]. H2 can be obtained from lignocellulosic biomass by and results of the studies are summarised in Table S1. The majority
various conversion technologies such as gasification (at N750 °C) and of the studies has focused on improving the photocatalytic
pyrolysis (at 450–650 °C) which are energy intensive [13,14], or by performance by either developing new photocatalysts such as TiO2/
thermocatalytic reforming which is associated with the problems of cat- NiOx@Cg [20], TiO2–S–Ni [21], CdS/CdOx quantum dots (QDs) [16],
alyst deactivation and by-product (e.g. CO) production [15]. Conversely, cellulose@Pt/TiO2 [22] and NCNCNx with H2PtCl6 [23] or improving
H2 production via the photocatalytic reforming of biomass-derived lig- the solubility of cellulose by introducing SO2−
4 to facilitate the acid
nocellulosic substrates in aqueous phase is a more sustainable route catalysed hydrolysis of cellulose. This generates soluble glucose
than the others regarding the energy source (i.e. the solar light) which is beneficial to quench the photo-generated holes which
required for driving the conversion. were used to oxidise water to producing hydroxyl radicals, and solu-
⁎ Corresponding authors.
ble glucose hence offering protons by the attack of hydroxide radi-
E-mail addresses: c.hardacre@manchester.ac.uk (C. Hardacre), cals [21,24]. Accordingly, the photocatalytic activity of cellulose
xiaolei.fan@manchester.ac.uk (X. Fan). reforming was improved considerably with improved H2 production

https://doi.org/10.1016/j.cjche.2020.03.030
1004-9541/© 2020 The Chemical Industry and Engineering Society of China, and Chemical Industry Press Co., Ltd. All rights reserved.
L. Lan et al. / Chinese Journal of Chemical Engineering 28 (2020) 2084–2091 2085

rates of N1000 μmol·h−1·g−1 cat (as shown in Table S1). However, only CuKα emission lines from a generator operating at 40 keV and 40 mA.
a few studies were attempted to investigate the kinetics of the pho- XRD scans were performed at 2θ from 10° to 90° with a step size of
tocatalytic reforming of plants-derived substrates [25,26], specifi- 0.033°. Crystallite sizes (D, nm) of the anatase and rutile phases were
cally cellulose [27,28,29]. The effect of the irradiation time, calculated using the Scherrer equation (Eq. 1). Anatase-rutile propor-
concentration of catalyst and cellulose, system pH and water salinity tions were determined by considering the relative diffraction inten-
on H2 generation has been investigated in the cellulose-water sus- sities of the (101) at 25.3° and (110) features at 27.4° corresponding
pension over Pt/TiO2 catalyst [27]. The H2 yield was improved with to the anatase and rutile phase, respectively. TEM micrographs were
an increase in the irradiation time, concentration of catalyst and cel- obtained with an FEI Tecnai G2 F20 electron microscope operated at
lulose before reaching the plateau, while the water salinity showed 200 kV. Before the measurements, catalyst samples were dispersed
insignificant effect on the yield of H2. The findings from the system- in ethanol and sonicated for 10 min, then followed by drop-casting
atic investigation [23,27,28] on the cellulose-based photocatalytic and drying the particle suspension on a copper TEM grid. SEM micro-
system were beneficial to explore the optimum operating condition graphs of materials were obtained by FEI Quanta 250 FEG microscope
of the system. Therefore, in addition to the exploration of the opti- operating at 15 kV.
mum operation condition of the relevant cellulose photoreforming
systems (by adjusting the reaction conditions), studies on the mech- kd λ
D¼ ð1Þ
anism is also needed to understand the possible pathways of H2 gen- β cosθ
eration from the systems.
This current work presents an investigation into the photocata- where kd is the dimensionless shape factor (kd = 0.943); λ is the X-
lytic reforming of cellulose using Pt/m-TiO2 (m-TiO2 with 80% ana- ray wavelength (λ = 0.154 nm); β is the line broadening in radians
tase and 20% rutile phases) catalysts under UV irradiation in water at half the maximum intensity of the diffraction peak at θ.
solution. m-TiO2 was used as the support to make the relevant cata- The optical property of the catalyst and cellulose were characterised
lysts under study comparable with the reported catalysts based on by UV–visible diffuse reflectance spectroscopy (UV–Vis DRS, Shimadzu
P25 [24–27] [29], as shown in Table S1. A systematic study was per- UV-2600) in the spectral range of 200–800 nm. N2 adsorption–desorp-
formed to study the effect of Pt loading, concentration of catalyst and tion analysis was performed at − 196 °C using a Micromeritics 3Flex
cellulose in the reaction system, and reaction temperature on the Surface Characterisation Analyser to determine the specific surface
photocatalytic performance, aimed at (i) establishing the relevant area of the catalysts (based on the Brunauer–Emmett–Teller (BET)
optimum reaction condition for improving the H2 production from method) and the pore size distribution of the catalysts (using the
the catalytic photocatalytic reforming of cellulose in aqueous system Barrett–Joyner–Halenda (BJH) method). Inductively coupled plasma
and (ii) obtaining the mechanistic understanding of the system, spe- optical emission spectroscopy (ICP-OES, PlasmaQuant PQ 9000 Elite)
cifically, analysing the reaction orders to identify the steps for H2 was performed to determine the actual Pt loading on the catalysts. Be-
Production. fore ICP-OES analysis, the sample (25 mg) was mixed with 12 ml aqua
regia (9 ml HCl: 3 ml HNO3) and digested using an ETHOS EASY micro-
2. Experimental wave digester at 220 °C for 20 min.
Temperature-programmed reduction (TPR) analysis of the catalysts
2.1. Chemicals and materials (~20 mg for each analysis) was measured under the 5% H2/Ar flow at
40 cm3·min−1. Prior to TPR, the catalyst was pre-treated in pure oxygen
m-TiO2, i.e. TiO2 with the mixed phase of rutile and anatase, was pur- (at 40 cm3·min−1) at 300 °C for 1 h. TPR analysis started at the initial
chased from Aldrich® and used as the catalyst support. Chloroplatinic temperature of −80 °C, under the 5% H2/Ar flow, and the temperature
acid hexahydrate (H2PtCl6·6H2O, Honeywell Fluka®) was used as the was then increased from − 80 °C to 800 °C with a heating rate of 10
metal precursor for catalyst preparation. Microcrystalline cellulose °C·min−1. The consumption of H2 was monitored by a thermal conduc-
(Aldrich®) with a density of 0.6 g·cm−3 was used in the photocatalytic tivity detector (TCD) in the QuantaChrom ChemBET Pulsar TPR/TPD
reaction as the biomass source. Deionised water was obtained from the Analyser.
Direct-Q 3UV ultrapure water system (Millipore®).
2.4. Photocatalytic reforming of cellulose over Pt/TiO2 catalysts
2.2. Preparation of photocatalyst
The experimental rig for the photocatalytic reforming of cellulose
Platinum catalysts supported on m-TiO2 (Pt/m-TiO2) were prepared was composed of a jacketed flat-bottomed borosilicate glass flask (as
by wet impregnation. 2 g m-TiO2 support was dispersed in X ml shown in the Support Information, SI, Fig. S1). During the experiment,
deionised water in a beaker and stirred for 10 min. Then Y ml (X + Y a circulating bath was used to circulate the heat transfer fluid through
= 20 ml, X and Y were varied during the preparation to adjust the Pt the jacket of the flask to maintain a constant reaction temperature at
loading on the catalysts) of the metal precursor solution (Cmp = 40 °C, 60 °C or 80 °C, respectively. On the top of the reactor, there are
0.01 g·ml−1, H2PtCl6·6H2O dissolved in deionised water) was added two vent holes which can be used to sample gaseous products from
into the TiO2 slurry. The mixture was stirred at 60 °C for 4 h, and the the reaction. For a typical photocatalytic reforming reaction experiment,
resulting catalyst was dried at 150 °C for 2 h and calcined in air at 500 75 mg catalyst and 100 ml distilled water were placed in the reactor,
°C for 2 h. Pt/m-TiO2 catalysts with different theoretical metal loadings, together with 100 mg cellulose. The mixture was stirred for 30 min at
ranging from 0.08 wt% to 1.5 wt%, were prepared, named as 0.08-Pt/m- room temperature, and the reactant mixture was purged with argon
TiO2, 0.16-Pt/m-TiO2, 0.50-Pt/m-TiO2, 1.00-Pt/m-TiO2 and 1.50-Pt/m- (Ar) thoroughly for 1.5 h to remove the dissolved oxygen from the
TiO2, respectively. Prior to the catalytic tests, all Pt/m-TiO2 catalysts system. After Ar purging, the reactor was sealed and irradiated by a
were reduced in pure H2 at 200 °C for 30 min. UV-A lamp (365 nm, 2 × 8 W, Thistle Scientific®) for 5 h. The initial
and final gaseous products from the headspace of the reactor were sam-
2.3. Characterisation of materials pled and analysed by gas chromatography (Perkin Elmer Clarus 580 GC,
fitted with 2 m inline HayeSep DB 100/120 mesh columns followed by a
The structure of the catalysts and cellulose was characterised using 2 m ShinCarbon ST 100/120 mesh column equipped with the TCD and
powder X-ray diffraction (PXRD), transmission electron microscopy FID detectors). The average production rate of gas phase (rM) is
(TEM) and scanning electron microscopy (SEM). XRD patterns of mate- defined as the mole amount of the gas (M) generated per hour
rials were obtained by PANalytical X'Pert Pro X-ray diffractometer using (μmol·h−1).
2086 L. Lan et al. / Chinese Journal of Chemical Engineering 28 (2020) 2084–2091

3. Results and Discussion were calculated using the PXRD results, as presented in Table 1. Varia-
tion of the crystalline size of the supports, as well as the proportion of
3.1. Characterisation of materials the anatase and rutile phase in the support, was measured after Pt im-
pregnation, i.e. anatase crystalline size in the support increased slightly
The nature of crystalline phases in Pt impregnated m-TiO2 was stud- and anatase/rutile ratio decreased. Considering the variation of anatase/
ied by PXRD as shown in Fig. S2. The diffractograms show that the pris- rutile phase ratio in the commercial P25 TiO2 (i.e. between 4:1 and 9:1),
tine m-TiO2 has the typical diffraction peaks corresponding to anatase the ratio of the catalysts after Pt impregnation (with an average value of
and rutile phases at 2θ = 25.5° and 27.5°, respectively, and the positions 4.6:1) was in the range.
of diffraction peaks remained the same after the Pt impregnation, as ex- N2 physisorption analysis was performed on the m-TiO2 support and
pected. The diffraction peaks associated with Pt species (metal or ox- the prepared catalysts (with the relevant adsorption–desorption iso-
ides) were not detected by XRD possibly due to the dispersion of Pt as therms shown in Fig. S3), and the relevant BET data are shown in
small nanoparticles of b5 nm. The crystallite size of TiO2 supports and Table 1. The m-TiO2 had a specific surface area of 99 m2·g−1 (as
the relative proportion of the anatase and rutile phases in the supports shown in Fig. S4). Pt impregnation resulted in the decrease in the BET
area of the support, possibly due to the partial occupation of Pt nanopar-
ticles in the porous network of the TiO2 support. The accurate Pt loading
Table 1
of the catalysts by ICP analysis was summarised in Table 1, showing the
Properties of m-TiO2 support and photocatalysts according to BET, XRD and ICP analyses
comparable values to the theoretical loadings.
Support Pt loading amount of catalysts/% The morphology of the catalysts was characterised by TEM, as shown
m-TiO2 0.08 0.16 0.50 1.00 1.50 in Fig. 1 and Fig. S5. Uniformly distributed Pt nanoparticles can be iden-
BET surface area/m2·g−1 99 55 42 53 53 58 tified for the catalysts with the relatively high loadings (i.e. N 1% shown
Anatase crystalline size/nm 16.5 19.6 20.2 19.5 18.8 18.7 in Fig. 1 c-d). For the catalysts with the Pt loading b1% (i.e. 0.16% and
Rutile crystalline size/nm 31.9 30.9 31.5 31.5 32.0 32.4 0.50% Pt catalysts supported on m-TiO2), Pt nanoparticles could not be
Anatase/rutile ratio 5.6:1 5.2:1 5.5:1 4.1:1 4.1:1 3.9:1 observed (shown in Fig. 1 a-b). The average Pt particle sizes for 1.00-
ICP Pt loading ratio/wt% 0 0.10 0.15 0.41 0.73 1.36
Pt/m-TiO2, 1.50-Pt/m-TiO2 are 1.7 nm and 1.8 nm, respectively.

Fig. 1. TEM images of (a) 0.16-Pt/m-TiO2, (b) 0.50-Pt/m-TiO2, (c) 1.00-Pt/m-TiO2 and (d) 1.50-Pt/m-TiO2.
L. Lan et al. / Chinese Journal of Chemical Engineering 28 (2020) 2084–2091 2087

The light absorption property of the support and catalysts was deter- catalysts, representing the reduction of PtOx crystallites to metallic Pt
mined by diffuse reflectance UV–Vis spectroscopy. Fig. S6 shows the [34,35].
absorbance spectra over the range between 200 and 800 nm for the
commercial m-TiO2 and Pt/m-TiO2 catalysts with different metal load- 3.2. Effects of Pt loading, catalyst concentration, cellulose concentration and
ings. The rapid increase in the absorption at wavelengths lower than reaction temperature on photocatalytic performance
380 nm can be assigned to the band gap absorption for m-TiO2 (band
gap energy: ~3.2 eV [30]). In the UV region (200–400 nm), Pt supported The effect of Pt loading in the catalysts on the production rate of H2
on m-TiO2 catalysts with Pt loadings from 0.16% to 1.5% show similar and CO2 was first investigated to compare the catalysts in the catalytic
absorbance intensities which are higher than that of m-TiO2 support photoreforming system (using cellulose as the substrate) for H2 gener-
and 0.08% Pt/m-TiO2. In the visible light region (400–800 nm), com- ation. As shown in Fig. 3a, the average production rate of H2 and CO2 (i.e.
pared with the m-TiO2 support, all the Pt/m-TiO2 catalysts show the rH 2 and rCO 2) over 5 h UV irradiation is plotted as a function of the Pt
absorption (as shown in insets in Fig. S6). Generally, the absorption of loading in the catalysts (H2 and CO2 were the main products measured
visible light by the catalysts increases with an increase in the Pt loading, in the gas phase, the overall gas composition of the systems is presented
which can be attributed to the localised surface plasmon resonance in Table S2). For the system using the m-TiO2 support, low rates of H2
absorption of Pt nanoparticles [31,32]. The optical absorption ability of were produced over 5 h reaction together with a small amount of CO2
cellulose was also investigated, as shown in Fig. S6. Generally, cellulose (rH 2 = 0.17 μmol·h−1, rCO 2 = 1.65 μmol·h−1). Pt/m-TiO2 catalysts sig-
has no capability of absorbing the visible light, while it shows the insig- nificantly promoted the rate of production of H2 and CO2, as shown in
nificant absorption of UV light (b 250 nm), which can be disregarded Fig. 3a, while the rH 2 and rCO 2 show the maxima as an increase of the
when compared with the UV light absorption of the catalysts. This result Pt loading. For Pt/m-TiO2 catalysts, the maximum rates were measured
indicates that the photochemical degradation of cellulose during the as about 9.95 and 3.75 μmol·h−1, respectively, for H2 and CO2, over the
catalytic photoreforming can be disregarded. 0.16-Pt/m-TiO2 catalyst. Pt species on TiO2 act as the electron traps [36],
The structure of the microcrystalline cellulose used in this study was promoting the separation of photo-generated electrons and holes, and
characterised by PXRD and SEM (as shown in Fig. 2). PXRD pattern of thus improving the photocatalytic activity in photoreforming reactions.
cellulose shows a profile with the typical peaks centred at 15.0°, 16.3° Accordingly, an increase in the Pt loading on TiO2 leads to the initial
and 22.5°, respectively, representing the cellulose I structure [28]. The increase in the activity of the system. However, larger loadings of Pt
crystallinity index (CrI) of the microcrystalline cellulose used is about on the surface of TiO2 may also hinder the light irradiation, reducing
80% according to Eq. 2 the generation of electrons and holes in the catalyst support. Therefore,
for the Pt/m-TiO2 catalyst, there is a trade-off between the two effects
CrI ¼ ðI002 −I AM Þ=I002  100% ð2Þ incurred by supporting Pt species, leading to the existence of an opti-
mum Pt loading on TiO2 supports. Different from the most studies in
where I002 represents the maximum intensity of the (002) lattice dif- which the highest rates were observed at Pt loadings higher than 0.5%
fraction (2θ = 22.5°), and IAM represents the intensity of diffraction [27–29], the highest rates were observed at a Pt loading of only
for the amorphous phase at 2θ = 18°. 0.16 wt% in this study. The performance of the 0.16-Pt/m-TiO2 catalyst
CrI is deemed as an important structural parameter regarding the in the photoreforming of cellulose under UV irradiation, represented
conversion of cellulose since a relatively low degree of crystallinity by the normalised rH 2 (by the mass of catalyst) at 132.6 μmol·h−1·g−1 cat ,
may result in the relatively fast hydrolysis of cellulose, leading to a com- was comparable with the reported state-of-the-art catalysts (shown
paratively easy interaction between the catalyst and the hydrolysed in Table S1) such as 0.50 wt% Pt/TiO2 (Hombikat TiO2, anatase/rutile
parts from cellulose [33]. Accordingly, in comparison with other natural (a/r) ratio = 19:6, under the irradiation of UV light at room temperature
cellulose substrate, such as cotton cellulose fibre (CrI = 80%–100%), the in water solution, 40 μmol·h−1·g−1 cat ) [29] and 0.50 wt% Pt/TiO2 (Evonik
use of microcrystalline cellulose (with the CrI of 80%) in photoreforming P25 TiO2, a/r ratio = 41:9, under the irradiation of UV light in water so-
can be advantageous. SEM micrograph of microcrystalline cellulose is lution, 225 μmol·h−1·g−1 cat ) [27]. The operation conditions over the 0.16-
shown in Fig. 2b, showing that the average particle size of the cellulose Pt/m-TiO2 catalyst and those two 0.50 wt% Pt/TiO2 catalysts are similar,
is approximately 100 μm together with a fraction of smaller particles. which are relatively mild compared with the conditions over other cat-
In order to investigate the reducibility of the TiO2 supports and Pt/ alysts (under the irradiation of Xe lamp in basic or acid solution at high
TiO2 catalysts, hydrogen temperature-programmed reduction (H2- temperature (N60 °C), shown in Table S1). Therefore, the possible rea-
TPR) analysis was carried out (as shown in Fig. S7). A H2 consumption son behind the variation of the photocatalytic performance among
peak at around 80 °C was identified in the H2-TPR profiles for all the them could be the differences in TiO2 supports. The 0.16-Pt/m-TiO2

Fig. 2. Characterisation of microcrystalline cellulose: (a) XRD diffractogram and (b) SEM image of cellulose.
2088 L. Lan et al. / Chinese Journal of Chemical Engineering 28 (2020) 2084–2091

Fig. 3. Gas production rates as a function of (a) Pt loadings on m-TiO2 (conditions: 0.75 g·L−1 Pt/m-TiO2, 1 g·L−1 of cellulose, under the irradiation of 5 h UV-A lamp for 5 h at 40 °C),
(b) catalyst concentrations (conditions: 0.16-Pt/m-TiO2, 1 g·L−1 of cellulose, under the irradiation of UV-A lamp for 5 h at 40 °C), (c) cellulose concentrations (conditions: 0.75 g·L−1
0.16-Pt/m-TiO2, cellulose, under the irradiation of UV-A lamp for 5 h at 40 °C), and (d) reaction temperature (conditions: 0.16-Pt/m-TiO2, 1 g·L−1 of cellulose, under the irradiation of
UV-A lamp for 5 h at 40 °C, 60 °C and 80 °C).

catalyst was employed in the following studies to investigate the effect from the reaction between photogenerated electrons (e−) and H+ in
of other parameters (including the catalyst concentration, the concen- the system. For the cellulose photoreforming system under study, H+
tration of cellulose and reaction temperature) on the photocatalytic per- is produced by oxidation of water and oxidation of cellulose. CO2 is pro-
formance since it gave the comparatively best performance during the duced from the oxidation of cellulose by hydroxyl radical (·OH) which
initial activity tests. is produced by H+ induced water oxidation. Once ·OH radicals are gen-
The concentration of 0.16-Pt/m-TiO2 catalyst was varied over range erated on the surface of the catalysts, they diffuse to the surface of the
of 0 to 1.5 g·L−1 to study its effect on the photocatalytic activity of the cellulose and are consumed by the cellulose. The consumption of ·OH
system (Fig. 3b). For the system with cellulose only (i.e. catalysts radicals increase with increasing the concentration of cellulose, how-
concentration = 0 g·L−1), CO2 and H2 were not produced under UV ever, when ·OH radicals are consumed completely, excess cellulose
irradiation due to the insignificant UV absorption of cellulose (Fig. S6). results in no increase in the CO2 production, thus leading to the plateau
By increasing the catalyst concentration in the system, the rH 2 and rCO 2 of the rCO 2.
showed a maximum. A relatively high concentration of the catalyst cor- The overall reaction of cellulose with water over Pt/m-TiO2 catalyst
responds to the improved UV light absorption, leading to the enhanced can be understood as follows:
photocatalytic activity. The increase in rH 2 and rCO 2 peaked at 9.95
μmol·h−1 and 3.75 μmol·h−1, respectively, up to 0.75 g·L−1. Further in- ðC6 H10 O5 Þn þ 7n H2 O→6n CO2 þ 12n H2 : ð3Þ
creases in the catalyst concentration from 0.75 g·L−1 to 1.5 g·L−1
caused the decrease in the photocatalytic performance which can be at- It was operated in a batch system; therefore, the overall rate law for
tributed to the hindered light absorption [37,38]. At high concentrations the process can be expressed by rH2 as:
of the catalyst, the resulting turbidity impedes the light penetration into
the inner section of the reactor, thus decreasing the light absorption by r H2
−r ¼ ¼ kC a1 C bH2 O ð4Þ
the photocatalyst in the bulk solution. 12n
Fig. 3c presents the effect of the cellulose concentration on rH 2 and
or it can also be expressed by rCO 2 as:
rCO 2 over the 0.16-Pt/m-TiO2 catalyst under UV irradiation. The photo-
catalytic activity was measured as insignificant for the system without r CO2
cellulose. By increasing the cellulose concentration in the reaction −r ¼ ¼ kC a2 C bH2 O ð5Þ
6n
system from 0.125 to 2 g·L−1, the H2 production increased sharply
and then decreased slightly to 2 g·L−1. Conversely, rCO 2 of the system where r is the overall reaction rate, C is the concentration of cellulose, a
plateaued at around 3.7 μmol·h−1 at 0.5 g·L−1 cellulose. H2 is produced (a1 for H2 and a2 for CO2) and b are the reaction orders with respect to
L. Lan et al. / Chinese Journal of Chemical Engineering 28 (2020) 2084–2091 2089

cellulose and water, respectively, n is the corresponding degree of poly- between the H+ derived from the oxidation of R-CHO into R-COOH
merisation of cellulose, and k is the rate constant. As the water in system and the photogenerated e−.
is excessive,CH 2O is considered unchanged, the Eq. 4 and Eq. 5 could be The thermal effect (40, 60 and 80 °C, respectively) on the photocat-
simplified as: alytic performance of the reaction system was also investigated, and the
result is illustrated in Fig. 3d. It is observed that an increase in the reac-
r H2 ¼ 2kobs C a1 ð6Þ tion temperature resulted in a decrease of rH 2 from 9.95 to 7.04
μmol·h−1, whereas it showed an insigfnicant influence on the rCO 2. By
r CO2 ¼ kobs C a2 ð7Þ increasing the solution temperature from 40 °C to 80 °C, the photo-
driven reaction steps cannot be affacted as the thermal energy kBT (kB
is Boltzmann constant = 8.617 × 10−5 eV·K−1, T in K) in the tempera-
where kobs = 6nkCbH2O. In order to compare a1 and a2, the logarithm of
ture range under study is too low (0.027 to 0.031 eV) to excite the semi-
Eq. (6) and Eq. (7) has been taken at both sides of the equation and
conductor (bandgap energy ≈ 3.2 eV) [40]. One plausible explanation of
the representation of ln (rH 2) (or ln (rCO 2)) versus ln (C) leads to a linear
the measured thermal effect on the activity (i.e. the reduced rH 2 due to
fit (shown in Fig. S8), and its slope corresponds to the reaction order a1
the temperature rise) has been propsoed based on the time-resolved
(or a2) with respect to cellulose:
spectroscopic study [41]. It was found that, once the electrons and
 holes in the semiconductor are excited by photons, either the recombi-
ln r H2 ¼ 1:661 ln ðC Þ þ 2:202;
nation or transfer to appropriate redox pairs is initiated, which is faster
  ð8Þ
R2 ¼ 0:979; 0:25bC g  L−1 b1 than diffusion and desorption of absorbed products on the catalyst.
Accoridngly, the production rate should be limited by “dark” catalytic
 steps, such as adsorption/desorption dynamics, diffusion of adsorbed
ln r CO2 ¼ 0:222 ln ðC Þ þ 1:556; species, etc. [26]. It could be concluded that the production of H2 is
  ð9Þ affected by the non-photo reaction steps which could be influenced by
R2 ¼ 0:971; 0:125bC g  L−1 b0:5 the thermal effect. However, CO2 is produced at the final step during
the non-photo reaction step, it could be influenced by the thermal effect
where the reaction order expressed by rH 2 (a1 = 1.661 in Eq. (8)) is but it is insignificant.
around 8 times higher than that of the reaction order expressed by Based on the systematic study of the catlytic photocatalytic
rCO 2 (a2 = 0.222 in Eq. (9)). reforming of cellulose, it was found that the investigated paremeters
The reaction order expressed by rCO 2 (a2 = 0.222) suggests that the indeed have influences on the performance of the photoreforming in
reaction is near-zero order with respect to cellulose, meaning that the aqueous system. Accoridng to the findings, the catalytic photoreforming
concentration of cellulose has a minor effect on rCO 2. Caravaca et al. system using cellulose is suggested being operated at 40 °C with
[28] suggested that photoreforming of cellulose I under the Xe lamp 1.0 g·L−1 of cellulose and 0.75 g·L−1 of 0.16-Pt/m-TiO2 catalyst to
irradiation was near zero-order, and this was also reported by Chang achieve the optimum photocatalytic performance (rH 2 = 9.95
et al. [29] when using both cellulose I polymorph or cellulose II poly- μmol·h−1 and rCO 2 = 3.75 μmol·h−1).
morph in the catalytic photoreforming system. However, in the latter Based on the finding from this study, in combination with the existing
studies, rH 2 was used solely to determine the reaction order, and on knowledge for the photoreforming of cellulose and glucose, the mecha-
the correlation between rCO 2 and the cellulose concentration is lacking. nism of H2 production from the catalytic photoreforming of cellulose
According to the previous research [28,37–39], during the process of was proposed, as shown in Fig. 4. The photocatalytic reaction is initiated
cellulose photoreforming, the hydrolysis of cellulose under the photo- by photons with energy (hv) equal to or higher than the bandgap energy
irradiation leads to the production of glucose, which was subsequently of the TiO2 support (Eg = 3.2 eV), leading to the generation of the e−-H+
oxidised by the •OH, giving aldehyde R-CHO such as glyceraldehyde pairs. Subsequently, the photo-generated e− and H+ transfer to the sur-
[38] and acid R-COOH such as formic acid [37,38]. Finally, H2 and CO2 face of the TiO2 to participate in redox reactions. This hypothesis is
were produced simultaneously via the oxidation of formic acid [39]. based on the fact that (i) the CB potential of TiO2 (around − 0.41 eV
Accordingly, based on this mechanism, i.e. the co-production of H2 and [42]) is more negative than the H+ reduction potential (0 V vs. the normal
CO2 via the oxidation reaction, the relevant reaction orders (i.e. a1 and hydrogen potential, HNE) and (ii) the VB potential of TiO2 (around
a2) should be comparable. Conversely, it was found at the first time 3.15 eV [43]) is more positive than the H2O oxidation potential (1.23 V
from the findings of the current kinetic analysis of the data, i.e. a1 ≈ 8 vs. NHE). Concerning the role of Pt species on TiO2, they act as the electron
a2, additional reaction routes exist in the photoreforming system in ad- sink to promote the separation of e− and H+ pairs, and hence avoiding
dition to the final step oxidation of formic acid, likely the reactions their recombination to some extents. Those photo-generated carriers

Fig. 4. Schematic model of the photocatalytic reforming of cellulose over Pt supported on TiO2 catalysts.
2090 L. Lan et al. / Chinese Journal of Chemical Engineering 28 (2020) 2084–2091

may promote the water-splitting reactions in the first place to create H+ research. YS thanks the CSC for her academic visiting fellowship at The
and ·OH from water (Fig. 4(a) process). The H+ and ·OH produced by the University of Manchester (file no. 201708440477) and the Foundation
oxidation of water then participate in the hydrolysis of cellulose, leading of Department of Education of Guangdong Province (No.
to the production of glucose (Fig. 4 (b) process). The ·OH radicals 2017KZDXM085, 2018KZDXM070).
would attack glucose, resulting in the generation of R-CHO and H+
(Fig. 4 (c) process), and the R-CHO would further be oxidised by ·OH rad- Supplementary Material
icals to R-COOH and H+ (Fig. 4 (d) process). Finally, the CO2 and H+
would be produced by the oxidation of R-COOH (Fig. 4 (e) process), and Supplementary data to this article can be found online at https://doi.
the H+ produced during the whole process would react with org/10.1016/j.cjche.2020.03.030.
photogenerated e− to produce H2 (Fig. 4 (f) process). During the whole
procedure, step (a) and (f) are the photo-driven reaction steps, affecting References
the production of H2, however, steps (b) to (e) are the non-photo reaction
[1] A. Onda, T. Ochi, K. Yanagisawa, Selective hydrolysis of cellulose into glucose over
steps which could be affected by thermal effect, influencing the produc-
solid acid catalysts, Green Chem. 10 (2008) 1033–1037.
tion of H2 as well as CO2. [2] Y. Wu, Z. Fu, D. Yin, Q. Xu, F. Liu, C. Lu, L. Mao, Microwave-assisted hydrolysis of crys-
talline cellulose catalyzed by biomass char sulfonic acids, Green Chem. 12 (2010) 696.
4. Conclusions [3] N. Yan, C. Zhao, C. Luo, P.J. Dyson, H. Liu, Y. Kou, One-step conversion of cellobiose to
C6-alcohols using a ruthenium nanocluster catalyst, J. Am. Chem. Soc. 128 (2006)
8714–8715.
In this work, a systematic study of the H2 production from the [4] S. Van de Vyver, J. Geboers, M. Dusselier, H. Schepers, T. Vosch, L. Zhang, G. Van
photocatalytic reforming of cellulose in aqueous systems over Pt cata- Tendeloo, P.A. Jacobs, B.F. Sels, Selective bifunctional catalytic conversion of cellu-
lose over reshaped Ni particles at the tip of carbon nanofibers, ChemSusChem 3
lysts supported on m-TiO2 (Pt/m-TiO2) was performed in order to un- (2010) 698–701.
derstand the photocatalytic behaviour of the system. The effect of Pt [5] L. Kong, G. Li, H. Wang, W. He, F. Ling, Hydrothermal catalytic conversion of biomass
loading (up to 1.50 wt%), catalyst concentration (0–1.5 g·L−1), cellulose for lactic acid production, J. Chem. Technol. Biotechnol. 83 (2008) 383–388.
[6] Y. Izumizaki, K.C. Park, Y. Tachibana, H. Tomiyasu, Y. Fujii, Organic decomposition in
concentration (0–2.0 g·L−1) and reaction temperature (40–80 °C) on supercritical water by an aid of ruthenium (iv) oxide as a catalyst-exploitation of
the photocatalytic activity of Pt/m-TiO2 was investigated to to identify biomass resources for hydrogen production, Prog. Nucl. Energ. 47 (2005) 544–552.
the optimum operation condition for the system, i.e. at 40 °C with [7] M. Osada, T. Sato, M. Watanabe, T. Adschiri, K. Arai, Low-temperature catalytic gas-
ification of lignin and cellulose with a ruthenium catalyst in supercritical water,
1.0 g·L−1 of cellulose and 0.75 g·L−1 over 0.16% Pt/m-TiO2 catalyst to
Energ. Fuel. 18 (2004) 327–333.
produce H2 at 9.95 μmol·h−1. The result suggested that even though [8] P.J. Dauenhauer, B.J. Dreyer, N.J. Degenstein, L.D. Schmidt, Millisecond reforming of
the cellulose photoreforming was operated in a relatively moderate solid biomass for sustainable fuels, Angew. Chem. Int. Ed. Engl. 46 (2007) 5864–5867.
[9] R. Lanza, D. Dalle Nogare, P. Canu, Gas phase chemistry in cellulose fast pyrolysis,
condition (UV-A lamp irradiation at 40 °C in pure water solution), a
Ind. Eng. Chem. Res. 48 (2009) 1391–1399.
low loading amount of Pt (0.16 wt%) on m-TiO2 support could result [10] S.E. Hosseini, M.A. Wahid, Hydrogen production from renewable and sustainable
in a sound average hydrogen production rate at 132.6 μmol·h−1·g−1. energy resources: Promising green energy carrier for clean development, Renew.
By comparing the reaction order a1 (expressed by rH 2) and a2 Sust. Energ. Rev. 57 (2016) 850–866.
[11] J.F. Madrid, L.V. Abad, Modification of microcrystalline cellulose by gamma
(expressed by rCO 2) with respect to cellulose and water, a1 is eight radiation-induced grafting, Radiat. Phys. Chem. 115 (2015) 143–147.
times as much as a2, suggesting the additional H2 production from the [12] N. Muradov, T. Veziroglu, “Green” path from fossil-based to hydrogen economy: An
oxidation reaction during the photoreforming process in addition to overview of carbon-neutral technologies, Int. J. of Hydrogen Energy 33 (2008)
6804–6839.
the co-production of H2/CO2 from photoreforming of cellulose in aque- [13] S. Turn, C. Kinoshita, Z. Zhang, D. Ishimura, J. Zhou, An experimental investigation of
ous system. Overall, the findings in this work established the optimum hydrogen production from biomass gasification, Int. J. Hydrog. Energy 23 (1998)
operation condition for photoreforming cellulose over the Pt/m-TiO2 641–648.
[14] S. Abdullah, S. Yusup, M.M. Ahmad, A. Ramli, L. Ismail, Thermogravimetry study on
catalyst, as well as sheding light on the addtional mechanism for H2 gen- pyrolysis of various lignocellulosic biomass for potential hydrogen production, Inter.
eration in the system. J. Chem. Biol. Eng. 3 (2010) 137–141.
[15] D. Wang, S. Czernik, E. Chornet, Production of hydrogen from biomass by catalytic
steam reforming of fast pyrolysis oils, Energ. Fuel. 12 (1998) 19–24.
Nomenclature
[16] D.W. Wakerley, M.F. Kuehnel, K.L. Orchard, K.H. Ly, T.E. Rosser, E. Reisner, Solar-
a, b Reaction orders with respect to cellulose and water driven reforming of lignocellulose to H2 with a CdS/CdOx photocatalyst, Nat. Energy
C Concentration of solution, g·l−1 2 (2017) 17021.
Cmp Metal precursor concentration, g·ml−1 [17] L.R. Lynd, P.J. Weimer, W.H. Van Zyl, I.S. Pretorius, Microbial cellulose utilization:
fundamentals and biotechnology, Microbiol. Mol. Biol. R. 66 (2002) 506–577.
CrI Crystallinity index, % [18] T. Kawai, T. Sakata, Conversion of carbohydrate into hydrogen fuel by a photocata-
D Crystallite size, nm lytic process, Nature 286 (1980) 474.
Eg Bandgap energy, eV [19] T. Kawai, T. Sakata, Photocatalytic hydrogen production from water by the decom-
position of poly-vinylchloride, protein, algae, dead insects, and excrement, Chem.
IAM Intensity of diffraction for the amorphous phase Lett. 10 (1981) 81–84.
I002 Maximum diffraction intensity of the (002) lattice [20] L. Zhang, W. Wang, S. Zeng, Y. Su, H. Hao, Enhanced H2 evolution from photocata-
k The rate constant in the overall rate law lytic cellulose conversion based on graphitic carbon layers on TiO2/NiOx, Green
0
kλ Chem. 13 (2018) 3008–3013, https://doi.org/10.1039/C8GC01398E.
k′ Dimensionless shape factor, k′ = 0.943 in D ¼
β cosθ [21] H. Hao, L. Zhang, W. Wang, S. Zeng, Facile modification of titania with nickel sulfide
kB Boltzmann constant = 8.617 × 10−5eV·K−1 and sulfate species for the photoreformation of cellulose into hydrogen,
ln (rM) Logarithm of the M production rate ChemSusChem 11 (2018) 2810–2817.
[22] G. Zhang, C. Ni, X. Huang, A. Welgamage, L.A. Lawton, P.K. Robertson, J.T. Irvine, Si-
n Corresponding degree of polymerisation of cellulose multaneous cellulose conversion and hydrogen production assisted by cellulose de-
r Overall reaction rate, μmol·h−1 composition under UV-light photocatalysis, Chem Commun (Camb) 52 (2016)
rM The production rate of a gas phase M, μmol·h−1 1673–1676.
[23] H. Kasap, D.S. Achilleos, A. Huang, E. Reisner, Photoreforming of lignocellulose into
β Line broadening at half the maximum intensity of the diffrac-
H2 using nanoengineered carbon nitride under benign conditions, J. Am. Chem. Soc.
tion peak at 2θ, rad 140 (2018) 11604–11607.
θ Diffraction angle, (°) [24] J. Zou, G. Zhang, X. Xu, One-pot photoreforming of cellulosic biomass waste to hy-
drogen by merging photocatalysis with acid hydrolysis, Appl. Catal. A Gen. 563
λ Wavelength of the light, nm
(2018) 73–79.
[25] D.I. Kondarides, V.M. Daskalaki, A. Patsoura, X.E. Verykios, Hydrogen production by
Acknowledgements photo-induced reforming of biomass components and derivatives at ambient condi-
tions, Catal. Lett. 122 (2008) 26–32.
[26] D.I. Kondarides, A. Patsoura, X.E. Verykios, Anaerobic photocatalytic oxidation of car-
LL thanks the China Scholarship Council (CSC, file no. 201706950035)- bohydrates in aqueous Pt/TiO2 suspensions with simultaneous production of hydro-
University of Manchester joint studentship for supporting her PhD gen, J. Adv. Oxid. Technol. 13 (2010) 116–123.
L. Lan et al. / Chinese Journal of Chemical Engineering 28 (2020) 2084–2091 2091

[27] A. Speltini, M. Sturini, D. Dondi, E. Annovazzi, F. Maraschi, V. Caratto, A. Profumo, A. [35] S. Kuhaudomlap, O. Mekasuwandumrong, P. Praserthdam, S.-I. Fujita, M. Arai, J.
Buttafava, Sunlight-promoted photocatalytic hydrogen gas evolution from water- Panpranot, The H2-treated TiO2 supported Pt catalysts prepared bystrong electro-
suspended cellulose: a systematic study, Photochem. Photobiol. Sci. 13 (2014) static adsorption for liquid-phase selective hydrogenation, Catalysts 8 (2018) 87.
1410–1419. [36] A. Naldoni, M. D’Arienzo, M. Altomare, M. Marelli, R. Scotti, F. Morazzoni, E. Selli, V.
[28] A. Caravaca, W. Jones, C. Hardacre, M. Bowker, H2 production by the photocatalytic Dal Santo, Pt and Au/TiO2 photocatalysts for methanol reforming: Role of metal
reforming of cellulose and raw biomass using Ni, Pd, Pt and Au on titania, Proc. Math. nanoparticles in tuning charge trapping properties and photoefficiency, Appl. Catal.
Phys. Eng. Sci. 472 (2016) 20160054. B Environ. 130 (2013) 239–248.
[29] C. Chang, N. Skillen, S. Nagarajan, K. Ralphs, J.T.S. Irvine, L. Lawton, P.K.J. Robertson, [37] P. Gomathisankar, D. Yamamoto, H. Katsumata, T. Suzuki, S. Kaneco, Photocatalytic
Using cellulose polymorphs for enhanced hydrogen production from photocatalytic hydrogen production with aid of simultaneous metal deposition using titanium di-
reforming, Sustain. Energ. Fuels. 3 (2019) 1971–1975. oxide from aqueous glucose solution, Int. J. Hydrog. Energy 38 (2013) 5517–5524.
[30] L. Qi, J. Yu, M. Jaroniec, Preparation and enhanced visible-light photocatalytic H2- [38] R. Chong, J. Li, Y. Ma, B. Zhang, H. Han, C. Li, Selective conversion of aqueous glucose to
production activity of CdS-sensitized Pt/TiO2 nanosheets with exposed (001) facets, value-added sugar aldose on TiO2-based photocatalysts, J. Catal. 314 (2014) 101–108.
Phys. Chem. Chem. Phys. 13 (2011) 8915–8923. [39] G. Wu, T. Chen, G. Zhou, X. Zong, C. Li, H2 production with low CO selectivity from
[31] Y. Zhu, D. Liu, M. Meng, H2 spillover enhanced hydrogenation capability of TiO2 used photocatalytic reforming of glucose on metal/TiO2 catalysts, Sci. China, Ser. B:
for photocatalytic splitting of water: a traditional phenomenon for new applications, Chem. 51 (2008) 97–100.
Chem. Commun. 50 (2014) 6049–6051. [40] V.M. Daskalaki, D.I. Kondarides, Efficient production of hydrogen by photo-induced
[32] J. Lu, L. Lan, X.T. Liu, N. Wang, X. Fan, Plasmonic Au nanoparticles supported on both reforming of glycerol at ambient conditions, Catal. Today 144 (2009) 75–80.
side of TiO2 hollow spheres for maximising photocatalytic activity under visible [41] A.V. Puga, Photocatalytic production of hydrogen from biomass-derived feedstocks,
light, Front. Chem. Sci. Eng. 13 (2019) 665–671. Coord. Chem. Rev. 315 (2016) 1–66.
[33] P.L. Dhepe, A. Fukuoka, Cellulose conversion under heterogeneous catalysis, [42] Y. Bessekhouad, D. Robert, J.V. Weber, Photocatalytic activity of Cu2O/TiO2, Bi2O3/
ChemSusChem 1 (2008) 969–975. TiO2 and ZnMn2O4/TiO2 heterojunctions, Catal. Today 101 (2005) 315–321.
[34] C. Zhang, H. He, K.-i. Tanaka, Catalytic performance and mechanism of a Pt/TiO2 cat- [43] V. Pfeifer, P. Erhart, S. Li, K. Rachut, J. Morasch, J. Brötz, P. Reckers, T. Mayer, S. Rühle,
alyst for the oxidation of formaldehyde at room temperature, Appl. Catal. B Environ. A. Zaban, I. Mora Seró, J. Bisquert, W. Jaegermann, A. Klein, Energy band alignment
65 (2006) 37–43. between anatase and rutile TiO2, J. Phys. Chem. Lett. 4 (2013) 4182–4187.

You might also like