Kageshima Et Al 2021 Insights Into The Electrocatalytic Oxidation of Cellulose in Solution Toward Applications in

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

pubs.acs.

org/JPCC Article

Insights into the Electrocatalytic Oxidation of Cellulose in Solution


toward Applications in Direct Cellulose Fuel Cells
Yosuke Kageshima,* Yurina Ojima, Hiromasa Wada, Sangho Koh, Masahiro Mizuno, Katsuya Teshima,
and Hiromasa Nishikiori*
Cite This: J. Phys. Chem. C 2021, 125, 14576−14582 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: The reaction mechanisms occurring during the electrocatalytic oxidation of


cellulose dissolved in a highly alkaline aqueous electrolyte were elucidated by hydrodynamic
voltammetry using a Pt rotating disk electrode as well as product analyses. The charge-
Downloaded via UNIV OF SYDNEY on March 10, 2024 at 02:31:29 (UTC).

transfer limited current associated with the multielectron process in which macromolecules
are cleaved into relatively short hydrocarbons was found to be dominant at relatively negative
potentials, whereas further decomposition of the hydrocarbons proceeded at more positive
potentials. Au was shown to facilitate the cleavage of cellulose macromolecules, while Pd and
Ni promoted additional oxidative decomposition of the short-chain hydrocarbons. A fuel cell
composed of Pt-deposited Ni foam electrodes as both the anode and cathode was capable of generating electricity in an external
circuit while directly utilizing cellulose as a fuel, even at ambient temperature and pressure. The present study provides new insights
into these reaction mechanisms and will assist in the design of catalytic materials intended for the effective utilization of biomass
energy sources.

■ INTRODUCTION
The effective utilization of the energy contained in biomass is
substrates5−7), the intrinsic activity of each metal has not yet
been established. Detailed insights regarding reaction mecha-
nisms and material properties will be vital to the design of
one of the most important next-generation technologies.
more effective catalytic materials so as to improve the
Cellulose, which accounts for the majority of waste biomass,
efficiency of DCFCs.
is an especially promising renewable energy source. Standard
On this basis, the present study examined the reaction
biomass reforming processes (such as thermal decomposition, mechanisms associated with the electrocatalytic oxidation of
saccharization, and fermentation)1−3 are labor-intensive and cellulose dissolved in strongly alkaline aqueous electrolytes
require significant energy expenditures, and so, the direct based on hydrodynamic voltammetry using a Pt rotating disk
utilization of cellulose as a fuel would be preferable. In fact, fuel electrode (RDE) as well as analyses of the resulting products.
cells based on inorganic electrocatalysts using cellulose as a fuel The electrocatalytic activities of several metal species intended
(known as direct cellulose fuel cells; DCFCs) have been to promote cellulose oxidation were also separately assessed
demonstrated.4−7 When these devices are operated at using disk electrodes.


sufficiently high temperatures (100−250 °C), solid cellulose
deposited on the heterogeneous electrocatalysts can be EXPERIMENTAL SECTION
decomposed to CO2 at a Faradaic efficiency of almost unity,
with the simultaneous generation of electricity in an external Electrolyte Preparation. A strongly alkaline aqueous
circuit.5 Cellulose can be dissolved in strongly alkaline aqueous solution containing cellulose was prepared by a freezing
solutions (such as 2 M NaOH)8−10 and thus can be used in method previously reported in the literature.6−10 In this
the form of a liquid fuel so as to provide a continuous reactant process, particulate microcrystalline cellulose (average molec-
supply to the DCFCs. It has been reported that low-molecular- ular weight: 40,000, average aspect ratio: 10, and diameter: 1−
10 μm) equivalent to 0.1−0.45 M glucose units (referred to
weight compounds such as glucuronic, oxalic, acetic, and
herein simply as 0.1−0.45 M cellulose) was dispersed in 2−8
formic acids are the major intermediate products during the
M NaOH solutions within perfluoroalkoxy alkane bottles, after
electrochemical decomposition of cellulose in solution.6
However, the details of associated reaction mechanisms
(such as the effect of the electrode potential on the reaction Received: May 17, 2021
pathway) are still unclear. Several electrocatalysts, including Pt, Revised: June 24, 2021
Au, Ni, and related alloys, have been used as anodes for Published: July 5, 2021
cellulose oxidation.5−7 Despite this, since these catalytic
materials have typically been assessed using porous electrodes
(such as by depositing metal catalysts on carbon black or foam

© 2021 American Chemical Society https://doi.org/10.1021/acs.jpcc.1c04357


14576 J. Phys. Chem. C 2021, 125, 14576−14582
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

which these dispersions were frozen by immersing the bottles


in liquid N2. After thawing the specimens at 40 °C, transparent,
viscous solutions were obtained. Solutions containing 2−8 M
NaOH without cellulose were made as controls.
Electrochemical Measurements. A commercially avail-
able Pt RDE or Pt, Au, Pd, Ni, or glassy carbon (GC) disk
electrodes were used as the working electrode. Before these
measurements, the disk electrodes were polished with
approximately 2 μm and 50 nm alumina particles. The trials
involving product analyses using a three-electrode config-
uration and demonstrations of DCFCs using a two-electrode
configuration employed Pt catalysts deposited on Ni foam
substrates (Pt/Ni) by sputtering acting as both the anode and
cathode.11 In the trials using a three-electrode configuration, a
commercially available reversible hydrogen electrode (RHE)
loaded with 2−8 M NaOH was added as the reference
electrode. The anolyte and catholyte were separated by a
cation-exchange membrane (ASTOM) in the two-chamber
cell.12 During the measurements, the anolyte and catholyte
were purged with Ar and O2, respectively. In the trials
involving product analyses, the electrochemical measurements
were conducted using a thin two-chamber cell structure
described in a previous paper12 to reduce the amount of each
electrolyte required and thus increase the concentrations of the
products.
Characterizations. Crystal structure analyses of the
particulate specimens were conducted by X-ray diffraction
(XRD; Rigaku, SmartLab) using Cu Kα radiation. In addition,
the pristine cellulose particles or the regenerated cellulose
specimens were pressed into KBr pellets and spectra were
obtained using Fourier-transform infrared (FTIR) spectrosco-
py (Shimadzu, IRPrestige-21). The water-soluble intermediate
products were identified using high-performance liquid
chromatography (HPLC) after each electrolysis trial.6,12 In
these experiments, residual cellulose was removed from the
electrolyte by first neutralizing the electrolyte with aqueous
HCl, after which the solid component was separated by Figure 1. Photographic images of (a) particulate cellulose dispersed in
centrifugation. The neutralized electrolyte was then filtered 2 M NaOH, (b) cellulose dispersion after freezing in liquid N2, (c)
through a hydrophilic nylon filter with a 0.22 μm pore size to 0.15 M cellulose in 2 M NaOH, (d) 0.1 M cellulose in 2 M NaOH,
(e) 0.3 M cellulose in 2 M NaOH, (f) 0.45 M cellulose in 2 M NaOH,
remove fine particles, and 20 μL of the specimen was injected (g) 0.15 M cellulose in 4 M NaOH, and (h) 0.15 M cellulose in 8 M
into an HPLC system (Shimadzu, Shimadzu Prominence LC- NaOH.
20A) equipped with an ultraviolet−visible detector (Shimadzu,
SPD-20A) and a Rezex RHM-Monosaccharide column
(Phenomenex, HPX-87H, 300 × 7.8 mm). During these increased. In particular, the aqueous solutions containing
analyses, the column temperature was maintained at 60 °C, 0.45 M cellulose in 2 M NaOH (Figure 1f) and 0.15 M
and an aqueous 4 mM H2SO4 solution was used as the eluent cellulose in 8 M NaOH (Figure 1h) gave highly viscous
at a flow rate of 0.8 mL min−1. The morphology of the Pt/Ni yellowish gel-like liquids. The evident changes in the color and
foam electrode was examined by scanning electron microscopy viscosity of the cellulose solution with increased concentrations
(SEM; Hitachi, SU8000).


can possibly be attributed to increased intermolecular
interactions.
RESULTS AND DISCUSSION The addition of aqueous HCl to these cellulose solutions
Characterizations for the Cellulose Dissolved in a resulted in a solid precipitate. The XRD patterns of the raw
Highly Alkaline Aqueous Solution. Various concentrations cellulose material showed characteristic peaks that could be
of particulate microcrystalline cellulose were dissolved in 2−8 assigned to crystalline cellulose I, whereas the material
M NaOH solutions using a previously reported freezing recrystallized from solution generated peaks assigned to
method.8−10 Changes in the appearance of the aqueous crystalline cellulose II (Figure 2a).8 These data indicate that
electrolyte during cellulose dissolution are shown in Figure a transition of the crystalline structure from cellulose I to II
1. In these trials, 0.15 M cellulose was dispersed in a 2 M occurred during the dissolution and subsequent regeneration
NaOH solution (Figure 1a), followed by freezing using liquid and that the structure of the cellulose skeleton was maintained
N2 (Figure 1b). After thawing the specimen at 40 °C, a through these processes. Both the pristine and regenerated
transparent, viscous solution was obtained (Figure 1c). With cellulose specimens also produced FTIR spectra containing
increases in the concentration of either cellulose (Figure 1c−f) absorption peaks that could be assigned to cellulose,8,13−17
or NaOH (Figure 1c,g,h), the viscosity of the solution providing further evidence for the retention of the primary
14577 https://doi.org/10.1021/acs.jpcc.1c04357
J. Phys. Chem. C 2021, 125, 14576−14582
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

intermolecular association between the cellulose chains, as well


as some degree of hydrolysis of the cellulose skeleton between
glucose units and the intramolecular decomposition of some
glucose derivatives. The amounts of these substances detected
by HPLC gradually increased with increasing NaOH
concentration, but no new compounds were observed. Thus,
the NaOH concentration affected the rate of hydrolysis rather
than the mechanism.
Assessments of the Electrocatalytic Oxidation of
Cellulose in Solution. A Pt electrode was found to generate
an anodic current resulting from cellulose oxidation over the
range of 0.9−1.5 VRHE (Figures S3−S5), and typical current−
potential curves obtained using a Pt RDE at various rotation
rates in highly alkaline electrolytes containing 0.15 M cellulose
are provided in Figure 3a. The results for the different
electrolyte concentrations are also summarized in Figure S6.
The anodic current at potentials more positive than 1.2 VRHE is
seen to increase with the increasing rotation rate, demonstrat-
ing the effect of diffusion as a limiting factor. Meanwhile, the
plateau current obtained at potentials of less than 1.2 VRHE is

Figure 2. (a) XRD patterns for the pristine particulate cellulose and
the regenerated cellulose obtained from a 2 M NaOH solution. (b)
FTIR spectra and (c) the corresponding crystallinity indices and
hydrogen-bond intensities (based on FTIR peak ratios as described in
the main text) for the cellulose recovered from NaOH solutions
having varying concentrations. (d) HPLC chromatograms of aqueous
solutions with varying NaOH concentrations after removing the
cellulose. Peak 1: solvent; peak 2: glucuronic acid; peak 3: malonic
acid; peak 4: succinic acid; peak 5: glycolic acid; and peak 6; formic
acid.

cellulose structure (Figure 2b). The details concerning the


assignment of each IR peak are presented in Figure S1 in the
Supporting Information. The absorption bands at 1420 and
890 cm−1 in these spectra were characteristic of crystalline and
amorphous structures, respectively, and the ratio of the
intensities of these peaks was defined as the “crystallinity
index”.14,18,19 It has also been reported that the ratio of the
absorbance in the range of 3000−3700 cm−1 to that at 990
cm−1 can be used as an empirical indicator of the “hydrogen-
bond intensity”, which in turn reflects the degree of
intermolecular regularity.14,20,21 Herein, the FTIR peaks at
890, 990, 1420, and 3000−3700 cm−1 are termed the
A(amorphous), A(CO), A(crystalline), and A(OH) peaks,
respectively, and the manner in which the peak heights were
determined is described in Figure S2. Dissolution of the
cellulose in the NaOH solutions obviously decreased both the
crystallinity index and hydrogen-bond intensity in the
regenerated cellulose compared with the original material,
although the concentration of NaOH had little effects on the
extent of these changes (Figure 2c). The decreased crystallinity
can be explained by the disruption of intermolecular
associations within the cellulose polymer domains by frozen
water, resulting in dissolution in the NaOH solution.10,22−24
HPLC analyses after the dissolution and precipitation of the Figure 3. (a) Current−potential curves generated using a Pt RDE
with various rotation rates in a 2 M NaOH solution containing 0.15
cellulose found small amounts of water-soluble organic acids in M cellulose. (b) Values for the degree of diffusion index nD2/3ν−1/6
the NaOH solution after the cellulose was removed.6,12 These and (c) Tafel plots of ik obtained in 2 M NaOH solutions containing
comprised oxidized derivatives of glucose (e.g., glucuronic various concentrations of cellulose as a function of electrode potential.
acid) as well as short-chain carboxylic acids (e.g., formic acid) (d) nD2/3ν−1/6 values and (e) Tafel plots obtained in 0.15 M cellulose
(Figure 2d). Therefore, the dissolution of cellulose in this solutions with varying concentrations of NaOH as a function of
highly alkaline solution appears to involve the cleavage of the electrode potential.

14578 https://doi.org/10.1021/acs.jpcc.1c04357
J. Phys. Chem. C 2021, 125, 14576−14582
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

almost insensitive to the rotation rate, implying that the cleaved from the cellulose at differing NaOH concentrations
contribution of the accelerated reactant diffusion to the overall (Figure 2d).
reaction rate (even in the case of rapid stirring) is almost To further clarify the reaction mechanisms, the effects of
negligible within this potential range. The reaction kinetics was electrode potential on the reaction products were elucidated
further analyzed using the Koutecky−Levich equation (Figure using a Pt/Ni foam11,12 as the anode (Figure 4a,b). In these
S7)
i−1 = ik −1 + (0.62nFD2/3ω1/2ν−1/6C)−1 = ik −1 + idiff −1
(1)
where ik, idiff, n, F, D, ω, ν, and C are the kinetically controlled
current (A cm−2), the limiting current for diffusion in the bulk
electrolyte (A cm−2), the number of electrons involved in the
reaction, Faraday’s constant (C mol−1 ), the diffusion
coefficient for the reactant (cm2 s−1), the rotation rate of the
disk electrode (rad s−1), the kinematic viscosity of the
electrolyte (cm2 s−1), and the bulk concentration of the
reactant (mol cm−3), respectively.25,26 Since the actual values
of n and ν in the present work were unclear and could be
affected by the electrolyte conditions, the term “nD2/3ν−1/6”
calculated from idiff = 0.62nFD2/3ω1/2ν−1/6C (i.e., the slope of
the Koutecky−Levich plots) was used as a measure of the
degree of diffusion in the electrolyte. In the case that diffusion
determines the overall reaction rate, this term should be
constant regardless of the electrode potential. Additionally,
more rapid diffusion of the reactant should increase the value
of this term, indicating a larger diffusion coefficient and/or
lower viscosity. The values of nD2/3ν−1/6 for 2 M NaOH
solutions containing different concentrations of cellulose are
plotted as a function of electrode potential (Figure 3b). At
positive potentials (1.3−1.5 VRHE), these values are independ-
ent of the electrode potential, demonstrating a diffusion-
limited process within this potential range. Regardless of the
electrode potential, higher cellulose concentrations resulted in
lower nD2/3ν−1/6 values, implying slower diffusion of the
reactants in the highly viscous solutions. The nD2/3ν−1/6 values
obtained in 0.1−0.15 M cellulose solutions also gradually
increase as the potential is decreased over the range of 1−1.2
VRHE. Because D and ν should be constant regardless of the
electrode potential for a given electrolyte, increases in this term Figure 4. (a,b) Current−potential curves obtained from Pt/Ni foam
likely reflect higher n, confirming a transition in the reaction in 2 M NaOH solutions with and without 0.15 M cellulose. (c)
pathway. Interestingly, the highly concentrated solutions (0.3− Current−time curves produced by Pt/Ni foam in a 0.15 M cellulose
0.45 M cellulose) did not show this change in the slope, solution at an applied potential of 1.1 or 1.5 VRHE. (d) Crystallinity
possibly due to the extremely small diffusion coefficient and/or index values and hydrogen-bond intensity for solid cellulose recovered
from the electrolyte after the electrocatalysis. (e) HPLC chromato-
high viscosity in such cases. The ik values, which represent the grams of the residual aqueous solution after removing cellulose from
electrocatalytic activities without the effects of mass transport the electrolyte following electrocatalysis. Peak 1: solvent; peak 2:
in the bulk electrolyte, were converted into Tafel plots (Figure oxalic acid; peak 3: glucuronic acid; peak 4: gluconic acid; peak 5:
3c), and the Tafel slope at more negative potentials up to 1 malonic acid; peak 6: succinic acid; peak 7: glycolic acid; peak 8:
VRHE was 114 mV decade−1 while that at potentials above 1.2 fumaric acid; and peak 9; formic acid.
VRHE was 209 mV decade−1. This change in slope also supports
switching of the elementary reaction mechanism over the
potential window of 1−1.2 VRHE, with the plateau current that trials, the residual cellulose and the electrolyte composition
was unaffected by the rotation rate (as seen in the current− were assessed using FTIR and HPLC, respectively, after
potential curves in Figure 3a) as a border. Because similar electrolysis. Although the anode generated the large redox
Tafel plots were obtained regardless of the cellulose current expected from Ni at approximately 1.3−1.4 VRHE, the
concentration, it appears that the cellulose concentration cellulose oxidation current was significantly enhanced
barely affected the underlying reaction process. The trials using compared with the value obtained with the flat Pt disk
0.15 M cellulose with different concentrations of NaOH also electrode due to the large surface area of the porous Ni foam
showed similar trends (Figure 3d,e). However, the Tafel slopes substrate (Figure S8). This increased surface area and the
at more positive potentials increased with increasing NaOH resulting increase in current would be expected to increase the
concentration despite the constant cellulose concentration concentration of various decomposition products in the
(Figure 3e), suggesting a change in the reaction kinetics as the solution and thus improve the qualitative analyses of these
NaOH concentration was varied. This result possibly reflects products by HPLC. In one experiment, cellulose solutions
changes in the concentration of the water-soluble compounds were electrolyzed by applying a constant potential of 1.5 or 1.1
14579 https://doi.org/10.1021/acs.jpcc.1c04357
J. Phys. Chem. C 2021, 125, 14576−14582
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

VRHE for several hours (Figure 4c), during which the anodic multielectron mechanisms. Interestingly, Au exhibited high
current was mainly attributed to cellulose oxidation (Figure activity at negative potentials in the range of 0.8−1.2 VRHE,
4a,b). The XRD patterns and FTIR spectra obtained from the while Pd and Ni showed much higher anodic currents
cellulose recovered after the electrolysis are presented in compared to the other metals at more positive potentials.
Figures S9 and S10. The crystallinity indices estimated from Considering the effect of the electrode potential on the
the FTIR spectra of the cellulose samples after electrolysis at reaction pathway as discussed above, it can be concluded that
1.5 or 1.1 VRHE were similar to those for cellulose that had Au promoted the cleavage of the cellulose macromolecular
simply been dissolved in NaOH without electrolysis, while the chains, while Pd and Ni drove further oxidative decomposition
hydrogen-bond intensities were obviously increased with of the short-chain hydrocarbons. Thus, alloying these elements
increasing applied potential (Figure 4d). These data indicate could be a potential approach to designing more active
that greater numbers of −OH groups were obtained as a result electrocatalysts for cellulose oxidation.
of the accelerated hydrolytic cleavage between the intra- Construction of the DCFC Using Pt/Ni Foam Electro-
molecular glucose units during electrolysis. The water-soluble des. Finally, a DCFC incorporating Pt/Ni foam electrodes
compounds produced during this process were analyzed by serving as both the anode and cathode connected in a two-
HPLC (Figure 4e), and the peak heights were found to have electrode configuration was demonstrated. The schematic
increased compared with the chromatogram obtained from drawing of the experimental setup and energy diagram during
cellulose simply dissolved in NaOH and then recovered the electric power generation are presented in Figure S14. The
without electrolysis. This observation indicates that the cyclic voltammogram acquired using the Pt/Ni foam during
hydrolysis of the cellulose skeleton and the decomposition of cellulose oxidation showed an onset potential for the anodic
the glucose derivatives were promoted by applying potentials current at a more negative potential than the onset potential
above 1.1 VRHE. When a more positive potential (i.e., 1.5 VRHE) for the cathodic current resulting from the oxygen reduction
was applied, the peaks assigned to low-molecular-weight reaction (Figure 6a), indicating that the present electro-
organic acids (e.g., formic acid) became greater, and a new
peak due to oxalic acid appeared, suggesting further progress of
the intramolecular decomposition. A kinetic analysis based on
the hydrodynamic voltammetry and the qualitative product
analyses showed that the charge-transfer limited current
associated with the extremely complex multielectron hydrolysis
reactions of macromolecular cellulose at the connections
between glucose units, and intramolecular cleavage to give
relatively short hydrocarbon chains, is dominant at negative
potentials in the range of 0.9−1.2 VRHE. In contrast, further
decomposition of the breakdown products involving relatively
small n values and in association with a reactant diffusion effect
proceeds at potentials above 1.2 VRHE.
The other metal disk electrodes also generated anodic
currents originating from the cellulose oxidation process
(Figures S11−S13), although it should be noted that the Au
and Ni electrodes additionally exhibited redox reactions of the
metals themselves.27,28 The effects of the electrode potential
on the cellulose oxidation current [expressed as the difference
between the anodic current in the cellulose solution (Icellulose)
and that in the NaOH solution (INaOH) as obtained from the
current−potential curves] when using Pt, Au, Pd, Ni, and GC
are summarized in Figure 5. The GC electrode produced a
negligible anodic current during cellulose oxidation, confirming Figure 6. (a) Current−potential curves acquired using Pt/Ni foam in
that the related reactions involved complicated inner-sphere, a 2 M NaOH solution containing 0.15 M cellulose (anolyte: red line)
and in 2 M NaOH saturated with O2 (catholyte: black line). (b) Time
course of the open-circuit voltage and (c) current−voltage profiles
obtained from a DCFC consisting of a Pt/Ni foam anode and
cathode. The results of control experiments without cellulose are also
compiled in (b,c).

chemical cell was capable of spontaneously generating


electricity in an external circuit. The electrocatalytic perform-
ances of the present Pt/Ni foam cathode in a 2 M NaOH
electrolyte under Ar or O2 bubbling are presented in Figure
S15 as a comparison. Indeed, the DCFC generated a stable
open-circuit voltage (VOC) of approximately 0.12 V, while the
electrochemical cell without cellulose was incapable of
spontaneously generating voltage (Figure 6b). The negative
Figure 5. Electrode potential dependence of the cellulose oxidation voltage observed for the control experiment without cellulose
currents obtained using various disk electrodes. may represent the chemical bias between the anolyte purged
14580 https://doi.org/10.1021/acs.jpcc.1c04357
J. Phys. Chem. C 2021, 125, 14576−14582
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

with Ar and catholyte purged with O2, which were separated by lose and NaOH, SEM image of the Ni foam, XRD
a membrane. That is, the electrode potential of Pt/Ni in the patterns and FTIR spectra of particulate cellulose
electrolyte saturated with O2 located at more negative potential samples regenerated from the aqueous electrolyte after
than the case with Ar, and thus, the concentration difference of the electrolysis process, current−potential curves for the
O2 could not contribute to the electric power generation due various metal disk electrodes, and current−time curves
to the large overpotential for the redox of O2/H2O. The acquired from the DCFC at different operating voltages
current−voltage profiles of the DCFC recorded during a cyclic (PDF)


voltage scan and steady-state polarization are presented in
Figure 6c, while the current−time curves for the steady-state
polarization are shown in Figure S16. Although the cyclic AUTHOR INFORMATION
voltage scan exhibited significant hysteresis due to the large Corresponding Authors
double-layer capacitance of the Ni foam substrate, which had a Yosuke Kageshima − Department of Materials Chemistry,
high surface area, the results obtained from the steady-state Faculty of Engineering and Research Initiative for Supra-
polarization experiment showed a VOC of 0.12 V and a short- Materials (RISM), Shinshu University, Nagano 380-8553,
circuit current (ISC) of 0.051 mA cm−2. The corresponding Japan; orcid.org/0000-0003-4919-9447;
power density profile is also provided in Figure S17. The VOC Email: kage_ysk@shinshu-u.ac.jp
determined based on the current−voltage profiles was in good Hiromasa Nishikiori − Department of Materials Chemistry,
agreement with the voltage−time curve acquired using open- Faculty of Engineering and Research Initiative for Supra-
circuit conditions. The electrochemical cell containing 2 M Materials (RISM), Shinshu University, Nagano 380-8553,
NaOH without cellulose as an anolyte could not generate Japan; orcid.org/0000-0002-6398-310X;
electricity, also supporting that the chemical energy of the Email: nishiki@shinshu-u.ac.jp
cellulose was converted into electricity in the present DCFC
device. It can therefore be concluded that a DCFC composed Authors
of a Pt/Ni foam anode and cathode generated electricity in an Yurina Ojima − Department of Materials Chemistry, Faculty
external circuit based on directly consuming cellulose in of Engineering, Shinshu University, Nagano 380-8553, Japan
solution as a fuel, even at ambient temperature and pressure. Hiromasa Wada − Department of Materials Chemistry,


Faculty of Engineering, Shinshu University, Nagano 380-
CONCLUSIONS 8553, Japan
Sangho Koh − Department of Bioscience and Textile
In summary, hydrodynamic voltammetry and product analyses
Technology, Interdisciplinary Graduate School of Science and
established that the electrocatalytic oxidation of cellulose
Technology, Shinshu University, Nagano 380-8553, Japan;
involves mid-chain scission of the cellulose macromolecules at
Present Address: Department of Chemistry for Life
more negative potentials and the further decomposition of the
Sciences and Agriculture, Faculty of Life Sciences, Tokyo
products into low-molecular-weight hydrocarbons at more
University of Agriculture, 1-1-1 Sakuragaoka, Setagaya-ku,
positive potentials. Although the reaction rate for the oxidation
Tokyo 156-8502, Japan
of the short-chain hydrocarbons at positive potentials was
Masahiro Mizuno − Department of Materials Chemistry,
determined by both charge-transfer and diffusion effects, the
Faculty of Engineering, Shinshu University, Nagano 380-
mid-chain scission of the cellulose was affected primarily by the
8553, Japan
rate of charge transfer without a reactant diffusion effect due to
Katsuya Teshima − Department of Materials Chemistry,
the multielectron processes involved. The data also showed
Faculty of Engineering and Research Initiative for Supra-
that Au promoted mid-chain scission at more negative
Materials (RISM), Shinshu University, Nagano 380-8553,
potentials, while Pd and Ni were capable of efficiently driving
Japan; orcid.org/0000-0002-5784-5157
the oxidation of the short-chain hydrocarbons at more positive
potentials. Thus, the present study provides new insights into Complete contact information is available at:
the kinetics of the complicated multielectron oxidation https://pubs.acs.org/10.1021/acs.jpcc.1c04357
reactions of polymeric macromolecules based on the use of
electrochemical techniques and also elucidates the intrinsic Notes
activities of several metal catalysts with regard to cellulose The authors declare no competing financial interest.
oxidation.

■ ASSOCIATED CONTENT
■ ACKNOWLEDGMENTS
This study was financially supported by a Grant-in-Aid for
*
sı Supporting Information Early-Career Scientists (no. 19K15675) from the Japan Society
The Supporting Information is available free of charge at for the Promotion of Science (JSPS).
https://pubs.acs.org/doi/10.1021/acs.jpcc.1c04357.
FTIR spectra obtained from particulate cellulose
regenerated from NaOH solutions having various
■ REFERENCES
(1) Nunes, L. J. R.; Causer, T. P.; Ciolkosz, D. Biomass for Energy:
concentrations and showing the manner in which the A Review on Supply Chain Management Models. Renewable
FTIR peak heights were determined, current−potential Sustainable Energy Rev. 2020, 120, 109658.
curves for a Pt RDE and a Pt disk electrode in aqueous (2) Caputo, A. C.; Palumbo, M.; Pelagagge, P. M.; Scacchia, F.
Economics of Biomass Energy Utilization in Combustion and
electrolytes containing various concentrations of cellu- Gasification Plants: Effects of Logistic Variables. Biomass Bioenergy
lose and NaOH, Koutecky−Levich plots of data 2005, 28, 35−51.
acquired at different electrode potentials using aqueous (3) McKendry, P. Energy Production from Biomass (Part 1):
electrolytes containing various concentrations of cellu- Overview of Biomass. Bioresour. Technol. 2002, 83, 37−46.

14581 https://doi.org/10.1021/acs.jpcc.1c04357
J. Phys. Chem. C 2021, 125, 14576−14582
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

(4) Li, Y.; Nagao, M.; Kobayashi, K.; Jin, Y.; Hibino, T. A Cellulose Aqueous Sodium Hydroxide and Supermolecular Structure of Soft
Electrolysis Cell with Metal-Free Carbon Electrodes. Catalysts 2020, Wood Pulp during Steam Explosion. Br. Polym. J. 1990, 22, 121−128.
10, 106. (24) Yamashiki, T.; Matsui, T.; Saitoh, M.; Matsuda, Y.; Okajima,
(5) Hibino, T.; Kobayashi, K.; Lv, P.; Nagao, M.; Teranishi, S. High K.; Kamide, K.; Sawada, T. Characterization of Cellulose Treated by
Performance Anode for Direct Cellulosic Biomass Fuel Cells the Steam Explosion Method. Part 3: Effect of Crystal Forms of
Operating at Intermediate Temperatures. Bull. Chem. Soc. Jpn. 2017, Original Cellulose on Changes in Morphology, Degree of Polymer-
90, 1017−1026. ization, Solubility Supermolecular Structure by Steam Explosion. Br.
(6) Hao, M.; Liu, X.; Feng, M.; Zhang, P.; Wang, G. Generating Polym. J. 1990, 22, 201−212.
Power from Cellulose in an Alkaline Fuel Cell Enhanced by Methyl (25) Koffi, R. C.; Coutanceau, C.; Garnier, E.; Léger, J.-M.; Lamy, C.
Viologen as an Electron-Transfer Catalyst. J. Power Sources 2014, 251, Synthesis, Characterization and Electrocatalytic Behaviour of Non-
222−228. Alloyed PtCr Methanol Tolerant Nanoelectrocatalysts for the Oxygen
(7) Sugano, Y.; Vestergaard, M. d.; Yoshikawa, H.; Saito, M.; Reduction Reaction (ORR). Electrochim. Acta 2005, 50, 4117−4127.
Tamiya, E. Direct Electrochemical Oxidation of Cellulose: A (26) Coutanceau, C.; Croissant, M. J.; Napporn, T.; Lamy, C.
Cellulose-Based Fuel Cell System. Electroanalysis 2010, 22, 1688− Electrocatalytic Reduction of Dioxygen at Platinum Particles
1694. Dispersed in a Polyaniline Film. Electrochim. Acta 2000, 46, 579−588.
(8) Yan, L.; Gao, Z. Dissolving of Cellulose in PEG/NaOH Aqueous (27) Görlin, M.; Chernev, P.; Ferreira de Araújo, J.; Reier, T.; Dresp,
Solution. Cellulose 2008, 15, 789−796. S.; Paul, B.; Krähnert, R.; Dau, H.; Strasser, P. Oxygen Evolution
(9) Cai, J.; Zhang, L. Rapid Dissolution of Cellulose in LiOH/Urea Reaction Dynamics, Faradaic Charge Efficiency, and the Active Metal
and NaOH/Urea Aqueous Solutions. Macromol. Biosci. 2005, 5, 539− Redox States of Ni−Fe Oxide Water Splitting Electrocatalysts. J. Am.
Chem. Soc. 2016, 138, 5603−5614.
548.
(28) Hoogvliet, J. C.; Dijksma, M.; Kamp, B.; van Bennekom, W. P.
(10) Isogai, A.; Atalla, R. H. Dissolution of Cellulose in Aqueous
Electrochemical Pretreatment of Polycrystalline Gold Electrodes to
NaOH Solutions. Cellulose 1998, 5, 309−319.
Produce a Reproducible Surface Roughness for Self-Assembly: A
(11) Kageshima, Y.; Shinagawa, T.; Kuwata, T.; Nakata, J.;
Study in Phosphate Buffer pH 7.4. Anal. Chem. 2000, 72, 2016−2021.
Minegishi, T.; Takanabe, K.; Domen, K. A Miniature Solar Device
for Overall Water Splitting Consisting of Series-Connected Spherical
Silicon Solar Cells. Sci. Rep. 2016, 6, 24633.
(12) Kageshima, Y.; Yoshimura, T.; Koh, S.; Mizuno, M.; Teshima,
K.; Nishikiori, H. Photoelectrochemical Complete Decomposition of
Cellulose for Electric Power Generation. ChemCatChem 2021, 13,
1530−1537.
(13) Lan, W.; Liu, C.-F.; Sun, R.-C. Fractionation of Bagasse into
Cellulose, Hemicelluloses, and Lignin with Ionic Liquid Treatment
Followed by Alkaline Extraction. J. Agric. Food Chem. 2011, 59, 8691−
8701.
(14) Oh, S. Y.; Yoo, D. I.; Shin, Y.; Seo, G. FTIR Analysis of
Cellulose Treated with Sodium Hydroxide and Carbon Dioxide.
Carbohydr. Res. 2005, 340, 417−428.
(15) Schwanninger, M.; Rodrigues, J. C.; Pereira, H.; Hinterstoisser,
B. Effects of Short-Time Vibratory Ball Milling on the Shape of FT-IR
Spectra of Wood and Cellulose. Vib. Spectrosc. 2004, 36, 23−40.
(16) Olsson, A.-M.; Salmén, L. The Association of Water to
Cellulose and Hemicellulose in Paper Examined by FTIR Spectros-
copy. Carbohydr. Res. 2004, 339, 813−818.
(17) Proniewicz, L. M.; Paluszkiewicz, C.; Wesełucha-Birczyńska, A.;
Majcherczyk, H.; Barański, A.; Konieczna, A. FT-IR and FT-Raman
Study of Hydrothermally Degraded Cellulose. J. Mol. Struct. 2001,
596, 163−169.
(18) Krässig, H. A. Cellulose; Structure, Accessibility and Reactivity;
Gordon and Breach Science: Philadelphia, 1993.
(19) O’Connor, R. T.; DuPré, E. F.; Mitcham, D. Applications of
Infrared Absorption Spectroscopy to Investigations of Cotton and
Modified Cottons: Part I: Physical and Crystalline Modifications and
Oxidation. Text. Res. J. 1958, 28, 382.
(20) Colom, X.; Carrillo, F. Crystallinity Changes in Lyocell and
Viscose-Type Fibres by Caustic Treatment. Eur. Polym. J. 2002, 38,
2225−2230.
(21) Nada, A.-A. M. A.; Kamel, S.; El-Sakhawy, M. Thermal
Behaviour and Infrared Spectroscopy of Cellulose Carbamates. Polym.
Degrad. Stab. 2000, 70, 347−355.
(22) Yamashiki, T.; Matsui, T.; Saitoh, M.; Okajima, K.; Kamide, K.;
Sawada, T. Characterization of Cellulose Treated by the Steam
Explosion Method, Part 1: Influence of Cellulose Resources on
Changes in Morphology, Degree of Polymerization, Solubility and
Solid Structure. Br. Polym. J. 1990, 22, 73−83.
(23) Yamashiki, T.; Matsui, T.; Saitoh, M.; Okajima, K.; Kamide, K.;
Sawada, T. Characterization of Cellulose Treated by the Steam
Explosion Method. Part 2: Effect of Treatment Conditions on
Changes in Morphology, Degree of Polymerization, Solubility in

14582 https://doi.org/10.1021/acs.jpcc.1c04357
J. Phys. Chem. C 2021, 125, 14576−14582

You might also like