Hibino Et Al 2018 Efficient Hydrogen Production by Direct Electrolysis of Waste Biomass at Intermediate Temperatures

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Research Article

Cite This: ACS Sustainable Chem. Eng. 2018, 6, 9360−9368 pubs.acs.org/journal/ascecg

Efficient Hydrogen Production by Direct Electrolysis of Waste


Biomass at Intermediate Temperatures
Takashi Hibino,*,† Kazuyo Kobayashi,† Masaya Ito,† Qiang Ma,† Masahiro Nagao,† Mai Fukui,‡
and Shinya Teranishi‡

Graduate School of Environmental Studies, Nagoya University, Nagoya 464-8601, Japan

Soken Inc., Nisshin, Aichi 470-0111, Japan
*
S Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: Biomass has been considered as an alternative


feedstock for energy and material supply. However, the lack of
high-efficiency and low-cost processes for biomass utilization and
Downloaded via UNIV OF SYDNEY on March 10, 2024 at 11:57:03 (UTC).

conversion hinders its large-scale application. This report


describes electrochemical hydrogen production from waste
biomass that does not require large amounts of energy or high
production costs. Hydrogen was produced by the electrolysis of
bread residue, cypress sawdust, and rice chaff at an onset cell
voltage of ca. 0.3 V, with high current efficiencies of approximately
100% for hydrogen production at the cathode and approximately
90% for carbon dioxide production at the anode. The hydrogen
yields per 1 mg of the raw material were 0.1−0.2 mg for all tested fuels. Electrolysis proceeded continuously at plateau voltages
that were proportional to the current. These characteristics were attributable to the high catalytic activity of the carbonyl-group
functionalized mesoporous carbon for the anode reaction, and that the major components of biomass such as cellulose, starch,
lignin, protein, and lipid were effectively utilized as fuels for hydrogen production.
KEYWORDS: Waste biomass, Electrolysis, Hydrogen production, Mesoporous carbon

■ INTRODUCTION
The Paris Agreement was adopted at COP 21 in 2015, so that
hydrogen. Furthermore, to obtain a high hydrogen yield, this
approach should be fully applied to all biomass components,
both developed and developing countries aim to reduce including starch, cellulose, hemicellulose, lignin, protein, and
greenhouse gas emissions by tens of percent by 2030.1−3 As fat.
part of this effort, the transition from fossil fuels to biofuels has Biomass electrolysis enables hydrogen production at onset
been accelerated, due to their much lower carbon footprint.4−6 cell voltages of less than 1 V, depending on the fuel species.
However, biofuels are not yet competitive with fossil fuels in Although bioethanol is one of the most promising fuels for
terms of cost and performance. To realize the commercializa- electrolysis,11,12 this fuel has the limitations inherent in the
tion of biofuels, various problems must be addressed, such as hydrolysis and fermentation of biomass, which requires special
the high total cost for biomass utilization, low efficiency of the and expensive procedures for processing. Biomass raw
conversion process into fuels, and insufficient reactivity of materials, including switchgrass and wood sawdust, can also
biomass raw materials.7,8 The use of waste biomass such as be used as fuels for electrolysis; however, polyoxometalates or
agricultural and food residues, harvested weeds, and wood chips ferric chloride are required to function as redox intermedi-
as feedstock decreases the raw material costs. Furthermore, if ates.13,14 The raw material previously reacts with the
hydrogen is directly produced from such waste biomass, then intermediate to form a reduced intermediate in a tank outside
processing costs and fuel availability could be significantly the electrolysis cell. The reduced intermediate is subsequently
improved. supplied to the anode and releases a proton to the cathode by
A major challenge for direct hydrogen production from the application of a voltage. We have recently reported another
biomass is how to gasify carbohydrates, proteins, and fats in the type of biomass electrolysis that does not require pretreatment
feedstock with high efficiency and low cost. The gasification of of the raw material.15 This method involves the addition (batch
biomass to hydrogen is generally conducted using water vapor operation) or supply (flow operation) of a mixed solution of
or air as a gasification agent at temperatures of 800 °C or newspaper and phosphoric acid (H3PO4) to the anode,
higher, followed by the separation and purification of
hydrogen,9,10 which is usually energy expensive. Therefore, it Received: April 15, 2018
is necessary to develop an alternative approach that leads to Revised: May 22, 2018
high-efficiency and cost-effective conversion of biomass into Published: May 24, 2018

© 2018 American Chemical Society 9360 DOI: 10.1021/acssuschemeng.8b01701


ACS Sustainable Chem. Eng. 2018, 6, 9360−9368
ACS Sustainable Chemistry & Engineering Research Article

Figure 1. Characterization of the unheated KB, and KB heated at 600 and 1000 °C: (a) TPD-MS spectra of CO and CO2 for the unheated KB. (b)
BJH pore size distributions. (c) FT-IR spectra. (d) XPS spectra.

followed by operation of the electrolysis cell at temperatures anode and as the cathode. Bread, cypress wood, and rice chaff were
between 100 and 150 °C. Cellulose16,17 and lignin18 are milled into powder form using a household mixer (Zojirushi BM-
hydrolyzed to mono- and disaccharide derivatives and aliphatic RT08-GA). Cellulose (Wako Chemicals), starch (Wako Chemicals),
molecules, respectively, by H3PO4 at the anode. These products lignin (Tokyo Chemical Industry), protein (Wako Chemicals), and
lipid (Wako Chemicals) were used as model biomass components
react with water (H2O) to form protons and carbon dioxide
without further purification.
(CO2) at the anode, and hydrogen at the cathode. For example, Characterization. Temperature-programmed desorption (TPD)
cellulose electrolysis is expected to proceed as follows. analysis for the oxygenated carbon sample was performed using online
mass spectrometry (MS; Pfeiffer Vacuum Thermostar GSD301). 0.5 g
Anode: C6H12O6 + 6H 2O → 6CO2 + 24H+ + 24e− of the sample powder was packed in a glass tube and supplied with Ar
(1) at a flow rate of 60 mL at 100 °C until the signals of carbon monoxide
(CO; m/z = 28) and CO2 (m/z = 44) were stable. The sample was
Cathode: 24H+ + 24e− → 12H 2 (2) subsequently heated to the desired temperature in an Ar flow at a rate
of 10 °C min−1. The pore characteristics of the carbon samples were
In this study, we present hydrogen production by the direct determined by nitrogen (N2) adsorption at liquid N2 temperature (Bel
electrolysis of waste biomass. First, the anode material is Japan BELSORP18PLUS-HT). The specific surface area was
modified to replace the expensive conventional Pt/C anode calculated using the Brunauer−Emmett−Teller (BET) method from
with metal-free mesoporous carbon. The electrolysis character- adsorption data in the relative pressure range of 0.001 to 0.3. The
istics of the pure biomass components are then evaluated using mesopore volume distributions were determined using the Barrett−
the optimized anode. Bread residue, cypress sawdust, and rice Joyner−Halenda (BJH) method. Functional group profiles for the
chaff are then examined as biomass feedstock in both batch- carbon samples were obtained using Fourier-transform infrared
spectroscopy (FT-IR; Agilent Excalibur FTS-4000). The chemical
and flow-type electrolysis cells. The hydrogen yield per weight
charge states of C 1s for the carbon samples were analyzed using X-ray
of biomass is determined in the batch cell. Continuous photoelectron spectroscopy (XPS; VG Escallab220i-XL). Morpho-
hydrogen production is demonstrated in the flow cell.


logical changes before and after heating the biomass raw materials
were observed using scanning electron microscopy (SEM; Keyence
MATERIALS AND METHODS VE-8800). The quantities of hydrogen and CO2 evolved from the
Materials. A Sn0.9In0.1P2O7-polytetrafluoroethylene (PTFE) mem- cathode and anode, respectively, were monitored using mass
brane was selected as the electrolyte, due to its high proton spectrometry (MS).
conductivity (>0.01 S cm−1) at temperatures between 100 and 150 Electrochemical Measurements. Electrolysis tests of batch and
°C, and a wide voltage range of 0−2 V.19−21 PTFE powder (0.04 g) flow cells were performed as follows (Figure S1). The anode (area: 1.1
was added to 1 g of Sn0.9In0.1P2O7 powder, kneaded using a mortar and cm2) was used for both cells. The cathode area was adjusted to 0.5 cm2
pestle, and then cold-rolled to a thickness of 170 μm using a laboratory for the batch cell and 0.2 cm2 for the flow cell, due to the difference in
rolling mill. Ketjen black (KB; EC-600JDK) was purchased from Akzo size between the cells. Details of the procedures have been described
Nobel and modified as follows. The carbon (1 g) was stirred in 50 mL previously.15 The current density was normalized with respect to the
of 24% nitric acid (HNO3) at room temperature for 74 h. After area of the cathode. In the batch cell, fuels (7.5−60.0 mg) were mixed
filtering and washing treatment, the residual carbon was dried under with appropriate quantities of 85% H3PO4 in a weight ratio of fuel to
vacuum at 120 °C for 6 h. The oxygenated carbon was subsequently 85% H3PO4 of 1:15. After deposition of the sample mixtures onto the
heated at 600 °C for 4 h in a flow of argon (Ar), unless otherwise surface of the anode, the electrolyte membrane was sandwiched
stated. The heat-treated carbon was dispersed with a small amount of between the anode and cathode. The anode was attached to a stainless
85% H3PO4 (Wako Chemicals) in a mixer (Thinky AR-100) for 30 steel current collector (SUS316, 16 mm diameter; Hohsen) and then
min. The slurry obtained was deposited on the surface of carbon fiber sealed using a PTFE gasket (Nitto Denko) with PTFE tape (Nitto
paper (Toray TGP-H-090). A commercially available Pt/C (Pt Denko). In the flow cell, a liquid mixture of fuel and 85% H3PO4 (fuel
loading: 2 mg cm−2, Electrochem) anode was used as a control concentration: 0.79 wt %) was pumped to the anode chamber with

9361 DOI: 10.1021/acssuschemeng.8b01701


ACS Sustainable Chem. Eng. 2018, 6, 9360−9368
ACS Sustainable Chemistry & Engineering Research Article

Figure 2. Electrolysis characteristics of the batch cells using the unheated KB anode, and the KB anodes heated at 600 and 1000 °C: (a) I−V curves
and (b) impedance spectra of the cells at 150 °C using the unheated KB anode, and the KB anodes heated at 600 and 1,000 °C. (c) I−V curves and
(d) impedance spectra of the cell using the 600 °C-heated KB anode between 100 and 150 °C. Data for the batch cell using the Pt/C anode are also
included.

flow channels at an injection rate of 16 mL h−1 using a syringe feeder in an Ar flow. Therefore, the oxygenated KB was heat-treated at
(YMC YPS-301). The anode chamber was sealed using the PTFE 600 and 1000 °C for test and control experiments, respectively.
gasket with an elastomeric O-ring. The cathode was supplied with Ar Figure 1b shows pore size distributions for the unheated KB,
at a flow rate of 100 mL min−1 in both cells. All electrochemical
measurements were conducted using a potentiostat-galvanostat
and the KB heated at 600 and 1000 °C. Similar characteristics
(Solartron 1287) and an impedance/gain-phase analyzer (Solartron were found for the three samples; the average mesopore
1260). Current−voltage (I−V) curves were recorded potentiostatically diameter and volume were 6.2, 7.1, and 6.8 nm and 2.6, 2.6, and
at a scan rate of 20 mV s−1. Voltage−time curves were acquired 2.7 cm3 g−1 for the unheated KB, and the KB heated at 600 and
galvanostatically in the current range of 0.05−0.20 A cm−2. Impedance 1000 °C, respectively. The BET specific surface areas for the
spectra were obtained at a bias voltage of 0.4 V in the frequency range samples were also almost unchanged among the samples, and
of 0.1−106 Hz. were 1864, 1772, and 1828 m2 g−1, respectively. These results

■ RESULTS AND DISCUSSION


Anode Design. For oxygen-functionalized carbon anodes,
demonstrate that even heat treatment at 1000 °C does not
cause significant structural deformation of the KB. As shown in
Figure 1c, the FT-IR spectrum for the HNO3-treated KB
the redox pair between carbonyl (CO) and phenol (C indicated the presence of carbonyl groups (absorption at 1720
OH) groups is one of the most important hydrogen carrier cm−1) and phenol groups (absorption at 1190 cm−1).28,29 The
sites.22,23 The origin of this redox reaction is ascribed to the intensities of these absorption bands decreased with an increase
following equilibrium reaction: of the heating temperature. To obtain additional information
regarding the effect of the heat treatment on the functional
CO + H+ + e− ⇌ COH (3) groups, XPS measurements were conducted for the KB before
In contrast, carboxyl groups (COOH) have little reversible and after the heat treatment. Figure 1d revealed less intensive
redox ability and are also one of the least conductive functional carbonylic CO (287.4 eV) and carboxylic COO (288.5 eV)
groups.24,25 Accordingly, the overpotential and resistive loss of peaks30 upon an increase of the heating temperature, the extent
the anode would be decreased by enhancement of the density of which was much greater for the COO peak (ca. 44% down)
of carbonyl groups relative to carboxyl groups on the carbon than for the CO peak (ca. 10% down). It was thus confirmed
surface. Figure 1a shows TPD profiles for CO (m/z = 28) and that the relatively carbonyl group-rich carbon sample was
CO2 (m/z = 44) in a flow of Ar with the HNO3-treated KB. successfully obtained by heat treatment at 600 °C, as expected
CO2 attributed to carboxyl group desorption26,27 was observed from the TPD profile.
in the temperature range of 150 to 550 °C, whereas CO Electrolysis of pure cellulose (15 mg) was conducted for the
originating from carbonyl group desorption26,27 became more unheated KB anode, and the KB anodes heated at 600 and
intensive at 550 °C or higher, which reflects the difference in 1000 °C. Figure 2a shows I−V curves for the electrolysis cells at
thermal stability between the two functional groups. This also a temperature of 150 °C. The onset cell voltage of electrolysis
implies that carboxyl groups are more preferentially eliminated was ca. 0.3 V for all the tested anodes, whereas the current
from the carbon surface than carbonyl groups by heat treatment density at each cell voltage increased in the order: 600 °C-
of the oxygenated KB at temperatures between 500 and 600 °C heated anode >1000 °C-heated anode > unheated anode.
9362 DOI: 10.1021/acssuschemeng.8b01701
ACS Sustainable Chem. Eng. 2018, 6, 9360−9368
ACS Sustainable Chemistry & Engineering Research Article

Figure 3. Electrolysis characteristics of the batch cell using various model biomass components as fuels at 150 °C: (a) I−V curves and (b) impedance
spectra. (c) Formation rates of hydrogen at the cathode, and CO2 and NO2 at the anode. Theoretical values are included for comparison. (d)
Formation rates of hydrogen, CO2, and NO2 during electrolysis of protein for 600 s at 0.25 A cm−2.

Impedance spectra for the electrolysis cells were measured at a transfer resistance obtained at 100 °C was too large for
bias voltage of 0.4 V (Figure 2b), and curve fitting was quantification. A similar temperature dependency of the
performed based on an equivalent circuit model (Figure S2). impedance was observed for the Pt/C anode; however, at all
The cell with the unheated anode presented a typical Nyquist temperatures, the semicircular arcs related to the mass-transfer
plot, the intersection of which with the abscissa at a high resistance were larger than those recorded for the optimized
frequency of 40,000 Hz was assigned to the ohmic resistance of mesoporous carbon anode (lower of Figure 2d). This is
the cell; the semicircular arcs at medium (3−40 000 Hz) and probably due to the difference in the mesopore volume
low (0.1−3 Hz) frequencies correspond to the polarization and between the carbon species used. Vulcan XC-72R used as the
mass-transfer resistances of the electrodes, respectively. This carbon support of the Pt/C anode has a total pore volume of
cell also possessed the largest ohmic and polarization 0.32 cm3 g−1,31 which is approximately one-eighth of that (2.6
resistances among the tested cells, which can be explained by cm3 g−1) observed for the KB anode. Therefore, the
the presence of large amounts of carboxyl groups introduced decomposition products of cellulose may not diffuse easily in
onto the surface through the HNO3 treatment. It is most likely such a small space.
that these functional groups lead to poor electrical contact Electrolysis of Pure Biomass Components. The
between the carbon particles. In contrast, the heat treatment of electrolysis of expected biomass components other than
such an oxygenated anode, especially at 600 °C, significantly cellulose was also investigated at a temperature of 150 °C.
decreased both the ohmic and polarization resistances due to Figure 3a shows I−V curves using various pure components
the removal of carboxyl groups without significant loss of (15 mg of each) as fuels. The current densities recorded with
carbonyl groups from the carbon surface. Subsequent the components at each voltage were one and 2 orders of
experimental trials were assessed using the 600 °C-heated magnitude greater than those obtained without fuel, which
anode because it was reasonable to conclude that there was no indicates that these components are viable electrolysis fuels.
optimal condition for the carbonyl group at temperatures above The electrolysis onset cell voltage was ca. 0.3 V for cellulose,
600 °C. starch, and protein and ca. 0.4 V for lignin and lipid. We cannot
Figure 2c shows I−V curves for the electrolysis cell using calculate the theoretical onset cell voltage for the electrolysis
cellulose as the fuel at temperatures between 100 and 150 °C. reactions, because these fuels are partially decomposed to
The I−V curve was dependent on the operation temperature, various products at the anode. However, as shown in Figure 2c,
which resulted in I−V characteristics recorded at 150 °C that the onset cell voltage using the 600 °C-heated KB and Pt/C
were comparable to the characteristics observed for the Pt/C anodes was independent of the anode species, but dependent
anode under the same conditions. This temperature depend- on the temperature. Therefore, it is most likely that the onset
ency was attributable mainly to the polarization and mass- cell voltage is determined by a thermodynamic factor rather
transfer resistances of the anode, rather than the ohmic than a kinetic factor. Another important result was the
resistance of the cell (upper panel of Figure 2d); the ohmic dependency of the I−V characteristics on the component
resistance remained almost unchanged and the polarization species: cellulose = starch > protein > lignin = lipid. The
resistance decreased from ca. 1.5 to ca. 0.3 Ω cm2 with an internal resistances of the cells with the component fuels were
increase in the temperature from 100 to 150 °C. The mass- compared using impedance measurements to better understand
9363 DOI: 10.1021/acssuschemeng.8b01701
ACS Sustainable Chem. Eng. 2018, 6, 9360−9368
ACS Sustainable Chemistry & Engineering Research Article

Figure 4. Photographs and SEM micrographs of (a, b, and c) bread, (d, e, and f) cypress, and (g, h, and i) rice chaff impregnated with H3PO4,
followed by heat treatment at 125 °C.

this dependency. As shown in Figure 3b, the impedance spectra four protons with two oxygen atoms in the fuel and H2O
revealed three distinctive features. First, similar polarization and molecules through the following four-electron reaction.
mass-transfer resistances were observed for cellulose (1.9 Ω
cm2) and starch (1.3 Ω cm2), which suggests a similarity of the C or N in fuel + 2O in fuel and H 2O
decomposition products (mono- and disaccharides) for these → CO2 or NO2 + 4H+ + 4e− (4)
components. Second, protein showed somewhat larger polar-
ization and mass-transfer resistances (6.2 Ω cm2) than those No products, except CO2 (and only NO2 for protein), were
acquired for cellulose and starch, which is attributable to the detected using MS. The current efficiency for CO2 formation
difference in the reducibility of the functional group in the was estimated to be 90% for cellulose, 85% for starch, 74% for
molecule between the carbonyl group in glucose and the lignin, 70% for lipid, and 67% for protein (see Supporting
carboxyl group in amino acid. Third, the polarization and mass- Information). It is likely that oxide compounds not detected by
transfer resistances obtained for lignin (92.7 Ω cm2) and lipid MS are also produced, especially from the lignin and lipid fuels.
(157.6 Ω cm2) were much larger than those recorded for the However, the current efficiency for NO2 formation from
other components. These results reflect that the two protein was 29%; therefore, almost all currents were considered
components are more acid-insoluble than the other compo- to be consumed for CO2 and NO2 formation at the anode.
nents, which leads to a reduction of available small molecule (The formation rates of these gases in addition to hydrogen
fuels at the anode. Figure 3b also showed that the ohmic during a short electrolysis period at 0.25 A cm−2 are shown in
resistance of the cell was especially large with the use of lignin Figure 3d). Based on these observations, it is concluded that
as a fuel. Approximately 60 wt % solid residue was observed hydrogen is produced by the electrolysis of most of the biomass
with the lignin fuel, which was identified as unreacted lignin, as components at onset cell voltages less than 0.5 V, which
previously reported.18 Therefore, the large ohmic resistance motivates the development of waste biomass electrolysis
observed for lignin is caused by poor electrical contact at the because these feedstock materials are composed mainly of the
interface between the anode and electrolyte, due to the biomass components tested.
presence of the solid residue. Electrolysis of Waste Biomass. Bread, cypress, and rice
Figure 3c shows hydrogen formation rates from the cathode chaff were tested as fuels in a similar manner to that for
at a temperature of 150 °C, including the theoretical values cellulose. The raw materials were used after being milled
(black dashed line) calculated according to Faraday’s law, based without purification; therefore, the species, concentration, and
on the two-electron reaction represented in Reaction 2. The fragment length of the biomass components varied according to
observed rates agreed approximately with the theoretical values the biomass source and the manufacturing process. This was
for all tested components, which demonstrates that the current confirmed by room temperature SEM observations, as shown in
efficiency for hydrogen production was 95−106% (see Figures 4b,e,h, where the sample powders were impregnated
Supporting Information). Figure 3c also shows CO2 and NO2 with H3PO4. The three raw materials presented intrinsic
formation rates from the anode. The theoretical rates shown by structures and sizes in the acid solvent. These samples were
the red dashed line in Figure 3c were calculated under the gradually liquefied when heated to 125 °C, but not completely
assumption that a carbon or nitrogen atom in the fuel is dissolved at this temperature (Figures 4c,f,i). In particular, rice
oxidized to a CO2 or nitrogen dioxide (NO2) molecule, and chaff usually contains ca. 15% silicon dioxide,32 which further
9364 DOI: 10.1021/acssuschemeng.8b01701
ACS Sustainable Chem. Eng. 2018, 6, 9360−9368
ACS Sustainable Chemistry & Engineering Research Article

Figure 5. Electrolysis characteristics of the batch cell using the bread, cypress, and rice chaff fuels at 150 °C: (a) voltage−time curves and (b)
formation rates of H2 at the cathode, and CO2 and NO2 at the anode during continuous electrolysis of the bread, cypress, and rice chaff (15 mg of
each) at 0.1 A cm−2. (c) Voltage−time curves during continuous electrolysis of rice chaff with various weights at 0.1 A cm−2. (d) Weight of produced
hydrogen as a function of the rice chaff fuel weight. Data for the bread and cypress fuels are also included.

Figure 6. Electrolysis characteristics of the flow cell using the bread, cypress, and rice chaff fuels at 150 °C: (a) voltage−time curves, (b) impedance
spectra, and (c) formation rate of hydrogen at the cathode during continuous electrolysis of the bread, cypress, and rice chaff at 0.1 A cm−2. (d)
Voltage−time curves and formation rate of hydrogen during continuous electrolysis of the bread at various current densities.

interferes with dissolution of the fragments into the acid materials: 70% starch in bread32 and 35% and 38% cellulose in
solvent. cypress33 and rice chaff,34 respectively. Furthermore, the
I−V curves using the biomass fuels were measured for batch magnitudes of the total impedance measured for the three
operation at a temperature of 150 °C (Figure S3a). Electrolysis fuels (Figure S3b) were closer to those obtained for cellulose
of the three fuels commonly began at cell voltages greater than and starch than those observed for protein, lignin, and lipid
ca. 0.3 V and provided current densities of ca. 0.3 A cm−2 at a (Figure 3b). Nevertheless, there were some differences in the
cell voltage of 1 V, which are characteristics similar to those ohmic, polarization and mass-transfer resistances among the
recorded for cellulose and starch. This is supported by the high cells with the three fuels, which is related to the differences in
content of starch or cellulose contained in these feedstock the component species and their proportions in the fuels.32−34
9365 DOI: 10.1021/acssuschemeng.8b01701
ACS Sustainable Chem. Eng. 2018, 6, 9360−9368
ACS Sustainable Chemistry & Engineering Research Article

The amount of available fuel at the anode decreased with the theoretical rates calculated using Faraday’s law. The
time for galvanostatic batch-mode electrolysis, which leads to electrical energy (kWh) consumed for electrolysis of each
the depletion of fuel at the anode at some point in time. fuel was calculated by integrating the area under the voltage−
Therefore, the quantity of produced hydrogen per unit weight time curve shown in Figure 6a and multiplying by the current.
of fuel can be estimated from the depletion time according to These energies were then normalized according to the volumes
Faraday’s law. Figure 5a,b shows voltage−time curves for the of experimentally produced hydrogen that were computed by
cells and formation rates for hydrogen, CO2, and NO2, integration of the data shown in Figure 6c. The estimated
respectively, during continuous electrolysis of the bread, energies were 0.75, 0.50, and 0.83 kWh (Nm3)−1 for bread,
cypress, and rice chaff (15 mg of each) fuels with a current cypress, and rice chaff, respectively, which are much less than
density of 0.1 A cm−2 at 150 °C. The voltage−time curves were those reported for water electrolysis [ca. 5 kWh (Nm3)−1] and
characterized by plateau-like behavior in voltage and by a ethanol electrolysis [ca. 2 kWh (Nm3)−1].35 It is evident that
subsequent rapid increase in voltage. These characteristics were these significant energy savings could be achieved by
associated with the consumption and depletion of the fuel suppression of the cell voltage to ca. 0.4 V at 0.1 A cm−2. To
because hydrogen and CO2 were constantly produced until the confirm such energy savings over a wide range of current
plateau-like behavior of the voltage was terminated. (The densities, electrolysis tests were performed using bread as the
current efficiencies for hydrogen, CO2, and NO2 were 98− fuel at various current densities. Figure 6d shows that the flow
101%, 83−99%, and 0−4%, respectively, during this period.) cell provided voltages of 0.19, 0.35, and 0.53 V at 0.05, 0.10,
The depletion time varied from 4000 to 6000 s, depending on and 0.15 A cm−2, respectively. There is a proportional
the fuel. Voltage−time curves obtained for rice chaff fuel with relationship between the voltage and current density, where
various weights are displayed in Figure 5c. The cell voltage the voltage increases by ca. 0.18 V per increment of 0.05 A
recorded at each time before depletion of the fuel became cm−2. However, at a current density of 0.20 A cm−2, the cell
gradually decreased as the fuel weight increased, which is voltage increased gradually from 0.64 to 0.72 V with time. The
reflected by an increase in the amount of the available fuel initial voltage of 0.64 V is likely counted lower, compared with
captured by the anode, as will be discussed later. More the other data sets, while the terminal voltage of 0.72 V is in the
importantly, the depletion time lengthened as the fuel weight proportional region. One possible explanation for this behavior
increased. As a result, the weight of the hydrogen estimated is the temperature increase of the cell due to the heat generated
from the depletion time increased with the weight of the fuel by the internal resistance of the cell, and subsequent recovery of
used, as shown in Figure 5d. Similar results were also obtained the temperature by an external heater. This also suggests the
for the bread and cypress fuels. Based on these results, 0.12− possibility of the thermal sustainability of the cell that would
0.22 mg, 0.14−0.24 mg, and 0.13−0.27 mg of hydrogen can be allow maintenance of the operation temperature in practical
produced from 1 mg of rice chaff, cypress, and bread, applications.36 Figure 6d also shows that hydrogen was
respectively. It is also noted that the hydrogen yield per unit produced in proportion to the current density and that the
weight of fuel tends to decrease with the weight of the fuel, due rate recorded at each current density was approximately 100%
to the deterioration of the utilization efficiency of the fuel at the of that calculated using Faraday’s law. These results, together
anode. This issue would be avoided by stirring the fuel in a with the results obtained for the batch cell, demonstrate that
batch cell. the biomass feedstocks examined could be electrolyzed
For batch-mode electrolysis, the voltage−time characteristics efficiently and continuously in one step to hydrogen. It is
could be improved by an increase in the weight of the fuel; also concluded that any of these feedstock materials could be
however, this effect decreased with time (Figure 5c), due to an utilized as fuels without significant differences in performance.
increase in the internal resistance of the cell. Accordingly, These characteristics are attributed to the effective production
continuous electrolysis characteristics of the present approach of hydrogen from most of the raw material components.
were reassessed in flow mode. Figure 6a shows voltage−time In conclusion, hydrogen could be successfully produced
curves for the flow cells with the bread, cypress, and rice chaff directly from biomass waste such as bread residue, cypress
fuels with a current density of 0.1 A cm−2 at 150 °C. Data sawdust, and rice chaff by electrolysis of a mixture of the raw
recorded for the batch cell with the rice chaff fuel (60 mg) is material and an 85% H3PO4 solvent at a temperature of 150 °C.
also included for comparison. The flow cells exhibited more A carbonyl-group functionalized mesoporous carbon exhibited
plateau-like voltages and lower cell voltages compared with the electrochemically catalytic activity for the anode reaction that
results obtained for the batch cells. The sudden small increase was comparable to the activity observed for a Pt/C anode. The
in voltage observed for the cypress fuel at ca. 2000 s is probably current efficiency for hydrogen production reached approx-
due to unknown disturbance of the flow in the fuel nozzle. The imately 100% for all the tested fuels in both the batch and flow
voltage plateau value at 3000 s was decreased in the order: rice operations. Note that the present approach provided more
chaff > bread > cypress. Figure 6b shows impedance spectra for significant savings in terms of the energy required for hydrogen
the flow cells at a bias voltage of 0.4 V, including data obtained production compared to that reported for the electrolysis of
for the batch cell with the rice chaff fuel. The difference in water and bioethanol.
impedance among the flow cells was too small to explain the
difference in the plateau voltage among them. However, note
that the impedances of the flow cells were much smaller than

*
ASSOCIATED CONTENT
S Supporting Information
that of the batch cell, which is ascribable to the constant supply
The Supporting Information is available free of charge on the
of fresh fuel to the anode in the flow process. ACS Publications website at DOI: 10.1021/acssusche-
Figure 6c shows the rates of hydrogen formation during meng.8b01701.
continuous flow electrolysis under the same conditions as those
shown in Figure 6a. The formation rate for hydrogen was Illustrations of electrolysis cells, impedance spectra for
almost the same in the three cells, which is in agreement with the cells at 150°C, and electrolysis characteristics of the
9366 DOI: 10.1021/acssuschemeng.8b01701
ACS Sustainable Chem. Eng. 2018, 6, 9360−9368
ACS Sustainable Chemistry & Engineering Research Article

batch cell using the bread, cypress, and rice chaff fuels at (12) Silveira, J. L.; Braga, L. B.; de Souza, A. C. C.; Antunes, J. S.;
150 °C (PDF) Zanzi, R. The benefits of ethanol use for hydrogen production in


urban transportation. Renewable Sustainable Energy Rev. 2009, 13,
2525−2534.
AUTHOR INFORMATION (13) Du, X.; Liu, W.; Zhang; Mulyadi, A.; Brittain, A.; Gong, J.;
Corresponding Author Deng, Y. Low-energy catalytic electrolysis for simultaneous hydrogen
*T. Hibino. E-mail: hibino@urban.env.nagoya-u.ac.jp. evolution and lignin depolymerization. ChemSusChem 2017, 10, 847−
854.
ORCID (14) Xu, F.; Li, H.; Liu, Y.; Jing, Q. Advanced redox flow fuel cell
Qiang Ma: 0000-0001-5993-819X using ferric chloride as main catalyst for complete conversion from
Author Contributions carbohydrates to electricity. Sci. Rep. 2017, 7, 5142.
(15) Hibino, T.; Kobayashi, K.; Ito, M.; Nagao, M.; Fukui, M.;
All authors contributed to the paper and approved the final Teranishi, S. Direct electrolysis of waste newspaper for sustainable
version. hydrogen production: an oxygen-functionalized porous carbon anode.
Funding Appl. Catal., B 2018, 231, 191−199.
Nagoya University received grants from the Japan Society for (16) Hibino, T.; Kobayashi, K.; Lv, P.; Nagao, M.; Teranishi, S.;
the Promotion of Science (JSPS): Nos. 26000008, 17H01895, Mori, T. An intermediate-temperature biomass fuel cell using wood
and 17K19087. sawdust and pulp directly as fuel. J. Electrochem. Soc. 2017, 164, F557−
F563.
Notes (17) Hibino, T.; Kobayashi, K.; Lv, P.; Nagao, M.; Teranishi, S. High
The authors declare no competing financial interest. performance anode for direct cellulosic biomass fuel cells operating at

■ ACKNOWLEDGMENTS
The authors thank Dr. Miki Niwa (Professor Emeritus, Tottori
intermediate temperatures. Bull. Chem. Soc. Jpn. 2017, 90, 1017−1026.
(18) Hibino, T.; Kobayashi, K.; Nagao, M.; Teranishi, S. Hydrogen
production by direct lignin electrolysis at intermediate temperatures.
ChemElectroChem 2017, 4, 3032−3036.
University) for numerous useful discussions. This work was (19) Lee, K. S.; Maurya, S.; Kim, Y. S.; Kreller, C. R.; Wilson, M. S.;
supported by Kakenhi Grants-in-Aid (Nos. 26000008, Larsen, D.; Elangovan, E.; Mukundan, R. Intermediate temperature
17H01895, and 17K19087) from JSPS.


fuel cells via an ion-pair coordinated polymer electrolyte. Energy
Environ. Sci. 2018, 11, 979−987.
REFERENCES (20) Hibino, T.; Kobayashi, K.; Nagao, M.; Kawasaki, S. High-
(1) Anderson, K.; Peters, G. The trouble with negative emissions. temperature supercapacitor with a proton-conducting metal pyrophos-
Science 2016, 354, 182−183. phate electrolyte. Sci. Rep. 2015, 5, 7903.
(2) Schellnhuber, H. J.; Rahmstorf, S.; Winkelmann, R. Why the right (21) Scott, K.; Xu, C.; Wu, X. Intermediate temperature proton-
climate target was agreed in Paris. Nat. Clim. Change 2016, 6, 649− conducting membrane electrolytes for fuel cells. Wiley Interdiscip. Rev.
653. Energy Environ. 2014, 3, 24−41.
(3) Aldy, J.; Pizer, W.; Tavoni, M.; Reis, L. A.; Akimoto, K.; Blanford, (22) Brunmark, A.; Cadenas, E. Redox and addition chemistry of
G.; Carraro, C.; Clarke, L. E.; Edmonds, J.; Iyer, G. C.; McJeon, H. C.; quinoid compounds and its biological implications. Free Radical Biol.
Richels, R.; Rose, S.; Sano, F. Economic tools to promote transparency Med. 1989, 7, 435−477.
and comparability in the Paris agreement. Nat. Clim. Change 2016, 6, (23) Revenga, J.; Rodriguez, F.; Tijero, J. Study of the redox behavior
1000−1004. of anthraquinone in aqueous medium. J. Electrochem. Soc. 1994, 141,
(4) Kumar, G.; Shobana, S.; Chen, W. H.; Bach, Q. V.; Kim, S. H.; 330−333.
Atabani, A. E.; Chang, J. S. A review of thermochemical conversion of (24) Shen, J.; Liu, A.; Tu, Y.; Foo, G.; Yeo, C.; Chan-Park, M. B.;
microalgal biomass for biofuels: chemistry and processes. Green Chem. Jiang, R.; Chen, Y. How carboxylic groups improve the performance of
2017, 19, 44−67. single-walled carbon nanotube electrochemical capacitors? Energy
(5) Gaurav, N.; Sivasankari, S.; Kiran, G.; Ninawe, A.; Selvin, J. Environ. Sci. 2011, 4, 4220−4229.
Utilization of bioresources for sustainable biofuels: a review. Renewable (25) Pantea, D.; Darmstadt, H.; Kaliaguine, S.; Sümmchen, L.; Roy,
Sustainable Energy Rev. 2017, 73, 205−214. C. Electrical conductivity of thermal carbon blacks: Influence of
(6) Voloshin, R. A.; Rodionova, M. V.; Zharmukhamedov, S. K.; surface chemistry. Carbon 2001, 39, 1147−1158.
Veziroglu, T. N.; Allakhverdiev, S. I. Review. Biofuel production from (26) Sebastián, D.; Suelves, I.; Moliner, R.; Lázaro, M. J. The effect of
plant and algal biomass. Int. J. Hydrogen Energy 2016, 41, 17257− the functionalization of carbon nanofibers on their electronic
17273. conductivity. Carbon 2010, 48, 4421−4431.
(7) Wainaina, S.; Horvath, I. S.; Taherzadeh, M. J. Biochemicals from (27) Figueiredo, J. L.; Pereira, M. F. R.; Freitas, M. M. A.; Ó rfão, J. J.
food waste and recalcitrant biomass via syngas fermentation: a review. M. Modification of the surface chemistry of activated carbons. Carbon
Bioresour. Technol. 2018, 248, 113−121. 1999, 37, 1379−1389.
(8) Arevalo-Gallegos, A.; Ahmad, Z.; Asgher, M.; Parra-Saldivar, R.; (28) Takaoka, M.; Yokokawa, H.; Takeda, N. The effect of treatment
Iqbal, H. M. N. Lignocellulose: a sustainable material to produce value- of activated carbon by H2O2 or HNO3 on the decomposition of
added products with a zero waste approach−a review. Int. J. Biol. pentachlorobenzene. Appl. Catal., B 2007, 74, 179−186.
Macromol. 2017, 99, 308−318. (29) Dandekar, A.; Baker, R. T. K.; Vannice, M. A. Characterization
(9) Sikarwar, V. S.; Zhao, M.; Fennell, P. S.; Shah, N.; Anthony, E. J. of activated carbon, graphitized carbon fibers and synthetic diamond
Progress in biofuel production from gasification. Prog. Energy Combust. powder using TPD and DRIFTS. Carbon 1998, 36, 1821−1831.
Sci. 2017, 61, 189−248. (30) Terzyk, A. P. The influence of activated carbon surface chemical
(10) Sikarwar, V. S.; Zhao, M.; Clough, P.; Yao, J.; Zhong, X.; composition on the adsorption of acetaminophen (paracetamol) in
Memon, M. Z.; Shah, N.; Anthony, E. J.; Fennell, P. S. An overview of vitro: Part II. TG, FTIR, and XPS analysis of carbons and the
advances in biomass gasification. Energy Environ. Sci. 2016, 9, 2939− temperature dependence of adsorption kinetics at the neutral pH.
2977. Colloids Surf., A 2001, 177, 23−45.
(11) Caravaca, A.; Sapountzi, F. M.; de Lucas-Consuegra, A.; Molina- (31) Chai, G. S.; Shin, I. S.; Yu, J. S. Synthesis of ordered, uniform,
Mora, C.; Dorado, F.; Valverde, J. L. Electrochemical reforming of macroporous carbons with mesoporous walls templated by aggregates
ethanol−water solutions for pure H2 production in a PEM electrolysis of polystyrene spheres and silica particles for use as catalyst supports in
cell. Int. J. Hydrogen Energy 2012, 37, 9504−9513. direct methanol fuel cells. Adv. Mater. 2004, 16, 2057−2061.

9367 DOI: 10.1021/acssuschemeng.8b01701


ACS Sustainable Chem. Eng. 2018, 6, 9360−9368
ACS Sustainable Chemistry & Engineering Research Article

(32) Nakamura, M.; Yuyama, Y. Development of a composition


database for various types of biomass. Technol. Rep. Natl. Res. Inst.
Agric. Eng. Japan 2005, 203, 57−80.
(33) Kawashima, A.; Akihiro, H.; Morita, H.; Fukuoka, M.; Honda,
K.; Morita, M. Dioxin-like polychlorinated biphenyl adsorbent
obtained from enzymatic saccharification residue of lignocellulose.
Bioresour. Technol. 2011, 102, 4682−4687.
(34) Johar, N.; Ahmad, I.; Dufresne, A. Extraction, preparation and
characterization of cellulose fibers and nanocrystals from rice husk. Ind.
Crops Prod. 2012, 37, 93−99.
(35) Lamy, C.; Jaubert, T.; Baranton, S.; Coutanceau, C. Clean
hydrogen generation through the electrocatalytic oxidation of ethanol
in a proton exchange membrane electrolysis cell (PEMEC): Effect of
the nature and structure of the catalytic anode. J. Power Sources 2014,
245, 927−936.
(36) Lotrič, A.; Sekavčnik, M.; Pohar, A.; Likozar, B.; Hočevar, S.
Conceptual design of an integrated thermally self-sustained methanol
steam reformer − High-temperature PEM fuel cell stack manportable
power generator. Int. J. Hydrogen Energy 2017, 42, 16700−16713.

9368 DOI: 10.1021/acssuschemeng.8b01701


ACS Sustainable Chem. Eng. 2018, 6, 9360−9368

You might also like