You are on page 1of 7

Article

pubs.acs.org/est

No Catalyst Addition and Highly Efficient Dissociation of H2O for the


Reduction of CO2 to Formic Acid with Mn
Lingyun Lyu, Xu Zeng, Jun Yun, Feng Wei, and Fangming Jin*
School of Environmental Science and Engineering, Shanghai Jiao Tong University, 800 Dongchuan Road, Shanghai 200240, People’s
Republic of China
*
S Supporting Information

ABSTRACT: The “greenhouse effect” caused by the


increasing atmospheric CO2 level is becoming extremely
serious, and thus, the reduction of CO2 emissions has become
an extensive, urgent, and long-term task. The dissociation of
water for CO2 reduction with solar energy is regarded as one
of the most promising methods for the sustainable develop-
ment of the environment and energy. However, a high solar-
to-fuel efficiency keeps a great challenge. In this work, the first
observation of a highly effective, highly selective, and robust
system of dissociating water for the reduction of carbon
dioxide (CO2) into formic acid with metallic manganese (Mn)
is reported. A considerably high formic acid yield of more than
75% on a carbon basis from NaHCO3 was achieved with 98%
selectivity in the presence of simple commercially available Mn powder without the addition of any catalyst, and the proposed
process is exothermic. Thus, this study may provide a promising method for the highly efficient dissociation of water for CO2
reduction by combining solar-driven thermochemistry with the reduction of MnO into Mn.

■ INTRODUCTION
The Earth’s surface temperature has risen by approximately
the redox of metals/metal oxides, such as Fe/Fe3O4, Zn/ZnO,
Mn(III)/Mn(II), Mn(IV)/Mn(II), and even MgxOy/Mg, using
0.85 °C from 1880 to 2012 according to the Intergovernmental solar energy have been achieved,17−19 and dissociation of water
Panel on Climate Change (IPPC 2013), which is a to hydrogen production20 was significantly higher than that
phenomenon caused by CO2 emissions from fossil carbon to with directive use of solar energy. Consequently, the
meet the energy demand of economic growth and civilization of dissociation of water with metal for the reduction of CO2
the human society.1−5 Presently, the concentration of CO2 is would represent one of the most promising approaches to
increasing markedly than any other time, in which senses, all increase artificial photosynthetic efficiency.
sectors, such as agriculture, food production, industry, tourism, Previous research has shown potential of the dissociation of
and health, are affected by this important phenomenon on the water for CO2 reduction with Fe. However, the yield of product
agenda.6−9 Melting of glaciers, rising of the water level in the formic acid was relatively low, even with the addition of a nickel
oceans, and vaporization in the fresh water resource as the heat catalyst, and the highest formic acid yield on a carbon basis was
increases all harm the natural balance and threaten the approximately 16%.21−23 Mn, as a first-row transition metal, has
ecological environment. an extraordinarily appealing coordination chemistry because of
A great deal of effort has been expended to reduce the CO2 its reactive redox nature.24 As a key element in photosynthesis,
concentration in the atmosphere, among which the solar Mn can mediate the splitting of water to provide the necessary
technologies are the ideal solution for the “greenhouse effect” electrons for photosynthesis.25−27 In addition, Mn plays a
problem.10−12 Artificial photosynthesis, in which solar energy is significant role in the synthesis of catalysts for CO2 hydro-
converted into chemical energy of renewable, non-polluting genation and Fischer−Tropsch (FT) synthesis.28 Moreover,
fuels and chemicals, is regarded as one of the most promising many researchers have reported a redox process of ZnO/Zn,
methods for the solar energy technologies. However, direct Mn(III)/Mn(II), and Mn(IV)/Mn(II) reactions using solar
conversion of solar energy into chemical energy retains many energy.18,19 Recently, Uchida et al. have also demonstrated that
problems, such as low conversion efficiencies and low product MgO/Mg can be circulated by solar power concentration using
selectivity. Developing an efficient solar-to-fuel conversion
process is a great and fascinating challenge.13−16 An integrated Received: November 22, 2013
system should be expected to improve artificial solar-to-fuel Revised: April 4, 2014
efficiency. Recently, some interesting integrated technologies of Accepted: April 18, 2014
a solar two-step water-splitting thermochemical cycle based on Published: April 18, 2014

© 2014 American Chemical Society 6003 dx.doi.org/10.1021/es405210d | Environ. Sci. Technol. 2014, 48, 6003−6009
Environmental Science & Technology Article

laser technology.17 The redox of MgO/Mg suggests that the used to investigate other possible chemicals in liquid samples.
reduction of MnxOy should be much easier than MgO using The system used a 2 mmol/L HClO4 solution as the mobile
solar energy because Mg is more active that Mn. Thus, a phase at a flow rate of 1.0 mL/min. The solid samples were
circulation of Mn should be achieved. Therefore, Mn may have washed with deionized water 3 times to remove impurities and
a much more significant implication than Fe in water splitting ethanol 3 times to make the solid sample quickly dry. The
for the conversion of CO2. samples were then dried in an isothermal oven at 40 °C for 3−5
Thus far, no study has reported the use of metallic Mn as an h and characterized using X-ray diffraction (XRD). XRD
efficient reductant of hydrogen production for the conversion analyses were performed on a Bruker D8 Advance X-ray
of CO2. With the goal of highly efficient dissociation of water diffractometer. The step scan covered angles of 10−80° (2θ) at
based on the redox of metals/metal oxides for the CO2 a rate of 2°/s.
reduction with a high yield, the use of Mn as a reductant to
produce hydrogen for CO2 reduction was investigated. We
found that Mn performed very well in the CO2 conversion
■ RESULTS AND DISCUSSION
Potential of CO2 Reduction with Mn. In a previous study
process. These new findings are reported in this paper. in which CO2 was reduced using Fe, the main product was

■ MATERIALS AND METHODS


Materials. Mn powder was obtained from Aladdin Chemical
formic acid.20 To investigate whether CO2 could be reduced to
useful chemicals or organics and to determine the reduction
products in the presence of Mn, experiments with NaHCO3
Reagent, and NaHCO3 was obtained from Sinopharm and Mn powder (200 mesh) were conducted at 300 °C for 2 h
Chemical Reagent Co., Ltd. In this study, NaHCO3 was used with a water filling of 35%, which is a good condition for
as a CO2 source to simplify handling. Gaseous CO2 and H2 obtaining a high yield of formic acid when using Fe as a
(>99.995%) were purchased from Shanghai Poly-Gas Tech- reductant. Liquid samples were analyzed by HPLC and GC/
nology Co., Ltd. Deionized water was used throughout the MS. As shown in Figure 1, HPLC analysis showed that the
study.
Experimental Procedure. Experiments were conducted
using a series of batch SUS 316 tubing reactors [9.525 mm (3/8
in.) outer diameter, 1 mm wall thickness, and 120 mm long]
with end fittings, providing an inner volume of 5.7 mL. Teflon-
lined reactors were used to examine the effect of the reactor
wall on the reaction for CO2 reduction. The schematic drawing
can be found elsewhere.29,30 The experimental procedure was
conducted as follows. The desired amounts of Mn, NaHCO3,
and deionized water were added to the reactor chamber. The
reactor was then sealed and put into a salt bath that had been
preheated to the desired temperature. After the preset reaction
time, the reactor was removed from the salt bath and then
placed into a cold water bath to quench the reaction. After
cooling to room temperature, the reaction mixture was
collected and filtered through a 0.22 μm syringe for analysis.
Figure 1. HPLC and GC/MS chromatogram of the liquid sample after
The water filling was defined as the ratio of the volume of the reaction (temperature, 300 °C; time, 2 h; NaHCO3, 1 mmol; Mn, 8
water put into the reactor to the inner volume of the reactor, mmol; water filling, 35%).
and the reaction time was defined as the duration of time that
the reactor was kept in the salt bath.
Product Analysis. The yield of formic acid was defined as main product was formic acid, and GC/MS analysis validated
the percentage of formic acid and the initial NaHCO3 on a this result, with very little acetic acid detected. Additionally, the
carbon basis as follows: TOC analysis showed that the selectivity of the production of
formic acid was over 98%, which was defined as the percentage
CF of formic acid and the TOC in liquid sample based on the
Y=
CS (1) carbon. Analysis of gas samples by GC/TCD showed that no
organic product was produced, and only hydrogen and a small
where CF and CS are the amounts of carbon in formic acid and amount of CO2 were detected. These results indicated that
in the initial NaHCO3 added to the reactor. Liquid samples formic acid was the main product from CO2 in the presence of
were filtered (0.22 μm filter film) and then analyzed by high- Mn. Quantitative analysis of the products obtained under this
performance liquid chromatography (HPLC), total organic condition showed that the formic acid yield was 43%. In
carbon (TOC), and gas chromatography/mass spectroscopy comparison to the reaction that used Fe as a reductant, in
(GC/MS). HPLC analysis was performed on KC-811 columns which the highest formic acid yield was only 16% with nickel as
(SHODEX) with an Agilent Technologies 1200 system, which a catalyst,20,22 the yield of formic acid in the presence of Mn
was equipped with a tunable ultraviolet/visible (UV/vis) was much higher. The results demonstrated that Mn was more
absorbance detector adjusted to 210 nm and a differential efficient in reducing CO 2 into formic acid than Fe.
refractometer detector. TOC was analyzed using a Shimadzu Subsequently, experiments with different sizes of Mn powder
TOC 5000A. Gas samples were analyzed by gas chromatog- were also conducted by changing the size of Mn powder from
raphy/thermal conductivity detector (GC/TCD). The Agilent 50 to 1400 mesh, and the results indicated that the formic acid
7890 GC/MS system, which was equipped with a 5985C inert yield has no evident change with different size Mn (see Figure
mass selective detector (MSD) and a triple-axis detector, was S1 of the Supporting Information). Thus, 200-mesh Mn
6004 dx.doi.org/10.1021/es405210d | Environ. Sci. Technol. 2014, 48, 6003−6009
Environmental Science & Technology Article

powder was chosen in this study. Further, the effect of the shown in eqs 3 and 4, and the decarboxylation of formic acid is
reactor wall of SUS 316 was also investigated with a Teflon- the predominant pathway in HTW.31,32
lined reactor, and the results demonstrated that no significant
HCOOH ↔ CO + H 2O (3)
catalytic role of the reactor wall of SUS 316 was observed (see
Table S1 of the Supporting Information).
In addition, an energy assessment for CO2 reduction into HCOOH ↔ CO2 + H 2 (4)
formic acid with Mn was also examined. According to the Thus, high reductant conditions may also inhibit the
proposed mechanism, the overall reaction could be described decomposition of the formed formic acid. The GC/TCD
by eq 2, and then the calculated reaction heat and free energy results confirmed this assumption, showing that only a small
using available thermodynamic date are negative. Thus, the amount of CO2 and no CO were present in the gas samples.
dissociation of water for CO2 reduction using Mn as reductant To examine the effect of the initial amount of NaHCO3
is not only spontaneous but also exothermic. It is known that, (carbon source) on the formic acid formation from CO2,
for an exothermic reaction, its equilibrium constant (Keq) will experiments were conducted at 300 °C for 2 h by fixing the Mn
decrease with an increase in the temperature. As expected, the amount at 8 mmol. As shown in Figure 2b, when the amount of
calculated Keq (600 K) is significantly lower than Keq (298 K) (the NaHCO3 was varied from 0.5 to 1.0 mmol, no obvious change
details can be found in the Supporting Information). on the yield of formic acid was observed. However, as the
Mn + CO2 + H 2O → MnO + HCOOH (2) amount of NaHCO3 further increased, the formic acid yield
decreased. No obvious change in the formic acid yield in the
° = −23.01 kJ/mol
ΔG298 ° = −114.75 kJ/mol
ΔH298 range between 0.5 and 1 mmol could be explained as more than
8:1 of the corresponding ratio of Mn/NaHCO3 because the
° = −10.78 kJ/mol
ΔG600 ° = −115.78 kJ/mol
ΔH600 formic acid yield remained nearly constant when the ratio of
Mn/NaHCO3 was above 8 mmol (see Figure 2a). For the
Keq (298 K) = e9.29 = 10829 Keq (600 K) = e 2.16 = 8.67 decrease in the formic acid with a further increase in NaHCO3,
Characteristics of the Reaction of Dissociation of it should be due to a decrease of the ratio of Mn/NaHCO3 to
Water for CO2 Reduction with Mn and the Parameter below 8 when NaHCO3 increased to above 1.0 mmol.
Design of the High Yield of Formic Acid. First, the effects Therefore, the reactant amounts of 8 mmol of Mn and 1
of the initial amounts of Mn and NaHCO3 were studied to mmol of NaHCO3 were chosen for the following study.
investigate characteristics of dissociation of water for CO2 The effect of the initial solution pH should be an important
reduction. As shown in Figure 2a, the initial amount of Mn factor in the reduction of CO2 to formic acid because pH can
affect the decomposition equilibrium of NaHCO3. The
decomposition of formic acid is also related to the pH of the
solution because alkaline conditions are generally not favorable
for the decomposition of formic acid.33 Takahashi et al.34 have
also reported that formic acid can be selectively formed by CO2
reduction in a weak alkaline solution under HTW conditions
when using Fe. Experiments to investigate the effect of pH were
conducted by adjusting the initial pH with NaOH or HCl. As
shown in Figure 3, the highest formic acid yield of 43%
occurred at the initial pH of 8.3, which is the same pH value as
that observed at 1 mmol of NaHCO3 with no additional NaOH
or HCl. In the cases of lower acidity, pH 6.6, and higher
alkalinity, pH 13.0, the yield of formic acid decreased to 34 and
10%, respectively. These results suggested that a weak alkaline
pH value of about 8.3 is favorable for the formation of formic

Figure 2. Effect of the amount of Mn and NaHCO3 on the formic acid


yield (temperature, 300 °C; time, 2 h; water filling, 35%; NaHCO3, 1
mmol for the effect of the amount Mn; Mn, 8 mmol for the effect of
NaHCO3).

strongly affected the formic acid yield at a fixed amount of


NaHCO3 (1 mmol). As the initial amount of Mn increased
from 2 to 10 mmol, the formic acid yield increased clearly from
13 to 43%, which was then remained nearly constant when the
amount of Mn was 8 mmol. The increase in the formic acid
yield with the increase in Mn most likely occurred because
stronger reduction conditions or a larger amount of hydrogen
improves the conversion of CO2. Additionally, it has been
reported that formic acid decomposes under high-temperature Figure 3. Effect of the pH on the formic acid yield (temperature, 300
water (HTW) via dehydration and/or decarboxylation, as °C; time, 2 h; NaHCO3, 1 mmol; Mn, 8 mmol; water filling, 35%).

6005 dx.doi.org/10.1021/es405210d | Environ. Sci. Technol. 2014, 48, 6003−6009


Environmental Science & Technology Article

acid, and thus, no additional alkali was used in the following In addition, considering that the increase in water filling lead to
experiments. a decrease in the initial concentration of NaHCO3, which may
Subsequently, the effects of the reaction temperature and increase costs for formic acid separation, further experiments
time on the conversion of CO2 were studied. As shown in were also conducted in the different water filling from 25 to
Figure 4, the yield of formic acid was very low and increased 55% by fixing the initial concentration of NaHCO3 and Mn as
the same as the optimal concentration (1 mmol of NaHCO3
and 8 mmol of Mn with 55% water filling) at 325 °C for 1 h. As
shown in Figure 5, the yield of formic acid increased clearly
with the increases in water filling when the ratio of NaHCO3/
Mn/H2O was constant, which showed that the increase in the
formic acid yield was directly related to the pressure but not the
NaHCO3 concentration.
CO2 Role of Improving Hydrogen Production from
Water. In our previous study, it was found that, in the absence
of NaHCO3, no hydrogen was produced when using Fe as the
reductant. However, a substantial quantity of hydrogen was
produced in the presence of NaHCO3, which indicated that an
increase in the initial NaHCO3 could lead to an increase in
hydrogen production. To determine whether NaHCO3 also
affects hydrogen production when using Mn as the reductant,
gas samples with and without NaHCO3 were collected and
Figure 4. Effect of the temperature and time on the formic acid yield analyzed by GC/TCD. As shown in Table 1, in the absence of
(NaHCO3, 1 mmol; Mn, 8 mmol; water filling, 35%).
Table 1. Amount of H2 and CO2 for Gas Samples and the
unconspicuously with the reaction time increasing at 250 and Yield of Formic Acida
275 °C. However, the formic acid yield increased obviously
when the temperature reached 300 °C, and the yield was as entry
Mn
(mmol)
NaHCO3
(mmol)
H2
(mL)
CO2
(mL)
yield of formic acid
(%)
high as 60% when the temperature reached 325 °C. Hence, a
1 4 0 80.5
high temperature was favorable for CO2 conversion. As shown
2 4 1 89.5 0.5 23
in Figure 4, the time profile indicated that the formic acid yield
3 8 1 140.0 43
increased rapidly at first, and a further increase in the reaction
time did not result in a significant increase in the yield of formic
a
At 300 °C for 120 min. The volume of total gas was measured at
acid. Therefore, 2 and 1 h were the optimal reaction times at room temperature of 20 ± 1 °C and pressure of 1 atm.
300 and 325 °C, respectively.
Finally, the effect of an important parameter of water filling
NaHCO3, 80.5 mL of hydrogen was produced when 4 mmol of
was investigated using constant initial amounts of NaHCO3 (1
Mn was used, whereas 89.5 mL of hydrogen was produced in
mmol) and Mn (8 mmol), with the water filling varying from
the presence of NaHCO3. Additionally, a formic acid yield of
25 to 55%. As shown in Figure 5, the formic acid yield
23% (0.23 mmol of formic acid) was obtained in the presence
of NaHCO3. Thus, the total hydrogen production in the
presence of NaHCO3 was higher than the total hydrogen
production without NaHCO3 because the formic acid formed
from CO2 hydrogenation consumed some amount of hydrogen.
Namely, CO2 can also promote hydrogen generation when
using Mn as a reductant, which is probably because the
oxidation of Mn shifts the reaction to the right because of the
consumption of hydrogen (CO2 hydrogenation) in the
presence of CO2, as shown in eq 5. Thus, CO2 provides an
additional benefit of improving hydrogen production from
water.

Possible Mechanism of Mn Oxidation in Water and


Role of MnxOy in the Formation of Formic Acid. The
Figure 5. Effect of the water filling on the yield of formic acid
mechanism of Mn oxidation was investigated by collecting the
(NaHCO3, 1 mmol; Mn, 8 mmol).
solid residues after reactions and analyzing them by XRD.
Interestingly, the rapid color change of the solid residues from
increased evidently with the increase in water filling, and the green to brown was observed during collection (see Figure 6a).
yields of 76 and 61% were attained as the water filling increased The color of MnO is green among oxides of Mn, and MnO is
to 55% at 325 and 300 °C, respectively. The results indicated easily oxidized to Mn3O4. Thus, Mn is most likely oxidized into
that the increase of water filling was favorable for CO2 MnO in the reactions, and the color change is attributed to the
reduction, which may be attributed to the increase in the further oxidation of MnO during collection in air. As expected,
pressure of the system because of the increase in water filling. XRD analysis showed that MnO and Mn3O4 were detected, as
6006 dx.doi.org/10.1021/es405210d | Environ. Sci. Technol. 2014, 48, 6003−6009
Environmental Science & Technology Article

Figure 6. (a) Photographs of solid samples and (b) XRD patterns of


the solid samples (NaHCO3, 1 mmol; time, 2 h; Mn, 8 mmol;
temperature, 300 °C; water filling, 35%).

shown in Figure 6b, and Mn(OH)2 was also formed. Thus,


there are two possible pathways for the oxidation of Mn to
MnO. One is that MnO is formed directly by the oxidation of
Mn, and the other is via the formation of Mn(OH)2, as shown Figure 7. (a) Photographs of solid samples and (b) XRD patterns of
in eq 6. the solid samples (NaHCO3, 1 mmol; Mn, 8 mmol; temperature, 300
°C; water filling, 35%).

To determine the reaction pathway of the oxidation of Mn into


MnO, experiments were conducted using short reaction times
of 1, 5, and 60 min. No green solid was observed during the
collection of solid samples, and XRD analysis showed that only
Mn(OH)2 was formed when the reaction times were l and 5
min, as shown in Figure 7. When the reaction time was
increased to 60 min, MnO and Mn3O4 were observed, whereas
the amount of Mn(OH)2 and Mn decreased gradually.
Simultaneously, the color change from green to brown in a
solid sample was observed during the collection of solid
samples. These results suggested that the oxidation of Mn to
MnO occurs via the formation of Mn(OH)2. Figure 8. Formic acid yields obtained with different amounts of
It is generally known that the use of a catalyst is needed for gaseous H2 and additive MnO or Mn3O4 (temperature, 325 °C; time,
activating hydrogen in the hydrogenation of CO2. However, 1 h; NaHCO3, 1 mmol; water filling, 55%; Mn, 8 mmol; gaseous H2, 6
interestingly, there is a high formic acid yield without the mmol; MnO, 6 mmol; Mn3O4, 6 mmol).
addition of any catalyst in the present study, which suggests
that some intermediates, such as MnxOy, formed in situ may act added (run 2); however, the formic acid was less than 2% (run
as a catalyst in the presence of Mn. To investigate this topic, 3) when Mn3O4 was added. The results indicated that MnO
experiments with NaHCO3 and gaseous hydrogen were can provide a catalytic activity in the reduction of CO2 into
conducted by changing the amount of gaseous hydrogen formic acid. However, the yield of 9% with MnO was much
from 5 to 18 mmol. As shown in Figure 8, all of the formic acid lower than that with Mn. One of the possible explanations is
yields with gaseous hydrogen were very low, keeping in only because MnO formed in situ when using Mn is more active than
about 2%, while a considerably high yield of formic acid can be the added MnO.
obtained when using Mn. These results suggested that MnxOy Investigation of Formic Acid Formation via HCO3− or
may act as a catalyst in the reduction of CO2 with Mn. To Gaseous CO2. After understanding the promotion of hydrogen
further provide evidence, experiments with gaseous hydrogen production in the presence of NaHCO3 and the mechanism of
and MnO or Mn3O4 additive were conducted. As shown in Mn oxidation, we also wanted to know whether the formation
Figure 8, the formic acid increased to 9% when MnO was of formic acid is via gas CO2 or HCO3−. To achieve this
6007 dx.doi.org/10.1021/es405210d | Environ. Sci. Technol. 2014, 48, 6003−6009
Environmental Science & Technology Article

objective, experiments with NaHCO3 and gaseous CO2 in the the Supporting Information, very little formic acid was
presence of Mn were conducted. As shown in Table S2 of the produced without water (entry 4), which indicated that water
Supporting Information, a formic acid yield of 43% was is necessary for the reduction of CO2 to formic acid in the
achieved with 1 mmol of NaHCO3 (entry 1). However, 0.2 and presence of Mn.
5.5% formic acid yields were produced, respectively, when using Furthermore, from a thermodynamic point of view, the
gaseous CO2 as a reactant instead of NaHCO3 in the absence of standard free energy change for CO2 hydrogenation with H
NaOH and in the presence of NaOH (entries 2 and 3), which into formic acid in the aqueous phase is −399.48 kJ/mol,
suggested that the formation of formic acid is closely related to whereas the reaction between H and CO2 in the gas phase is
the concentration of CO2 in the solvent, and the low −373.48 kJ/mol (eqs 9 and 10). The details were presented in
concentration of CO2 in the solvent is not favorable for the the Supporting Information.
formic acid formation. If this hypothesis is true, increasing the
dissolution time of CO2 before the reactions should lead to an ° = −399.48 kJ/mol
ΔG298
increase in the formic acid yield. To test this hypothesis, HCO3− + 2H ⇌ HCOO−
° = −411.35 kJ/mol
ΔH298
experiments were performed with an additional dissolution + H 2O
process at room temperature before the hydrothermal reaction (9)
in an attempt to increase the dissolution of CO2 in NaOH
solution. As expected, the yield of formic acid clearly increased ° = −373.48 kJ/mol
ΔG298
and the initial pH of solution decreased with the prolongation CO2 + 2H ⇌ HCOOH
° = −467.18 kJ/mol
ΔH298
of the dissolution time; the best formic acid yield was achieved
at a weak alkaline pH value, as shown in Figure 9. This (10)
The results indicated that the reaction is exothermic and
HCO3− is preferred over CO2 for the interconversion between
hydrogen and formic acid.
In conclusion, a novel method of no catalyst addition and
highly efficient dissociation of water for highly selective
conversion of CO2 into formic acid with Mn powder as a
reductant was developed. A considerably high formic acid yield
of more than 75% was obtained from CO2 with simple
commercially available Mn powder. CO2 not only acts as a
carbon source but also improves hydrogen production from
water. The present study is helpful for providing a promising
method for highly efficient dissociation of H2O for CO2
reduction combined with MnO/Mn using solar energy.


*
ASSOCIATED CONTENT
S Supporting Information
Figure 9. Yield of formic acid with the dissolution of CO2 (Mn, 8 Formic acid yield with a Teflon-lined reactor (Table S1),
mmol; CO2, 1 mmol; NaOH, 1 mmol; water filling, 35%; dissolution difference in the yields of formic acid between NaHCO3 and
of CO2, at room temperature; temperature, 300 °C; time, 2 h).
CO2 gases (Table S2), effect of the amount of metal reductant
ratio and Mn size on the formic acid yield (Figure S1), and
observation is in agreement with the fact that the best formic energy assessment for reduction of CO2 into formic acid with
acid yield was achieved using a weak alkaline pH. CO2 can exist Mn. This material is available free of charge via the Internet at
as different forms of hydrogen carbonate and carbonate at http://pubs.acs.org.
different pH values, as shown in eq 7, and the distribution
coefficient of HCO3− is more than 0.9 when the pH value is
8.3, as shown in eq 8.
■ AUTHOR INFORMATION
Corresponding Author
pK1= 6.3 + pK 2 = 10.3 *Telephone/Fax: +86-21-54742283. E-mail: fmjin@sjtu.edu.cn.
CO2 + H 2O XoooooooY HCO3− + H XoooooooooY CO3 2−
+ 2H +
Notes
(7)
The authors declare no competing financial interest.
pH = p
K 2[HCO3−] + KW
1+
[HCO3−]
K1
= p K 2K1
■ ACKNOWLEDGMENTS
The authors gratefully acknowledge financial support from the
1 National Natural Science Foundation of China (Grants
= (pK1 + pK 2) = 8.3 21077078 and 21277091).


2 (8)
As discussed previously, the highest formic acid yield occurred REFERENCES
at an initial pH of 8.3. The yield of formic acid significantly (1) Cox, P. M.; Betts, R. A.; Jones, C. D.; Spall, S. A.; Totterdell, I. J.
decreased under more acidic or more alkaline conditions, which Acceleration of global warming due to carbon-cycle feedbacks in a
may be attributed to the fact that the main ion present is not coupled climate model. Nature 2000, 408 (6809), 184−187.
HCO3− when the pH is less than 6.3 or more than 10.3. These (2) Ah-Hyung Alissa, P.; Liang-Shih, F. CO2 mineral sequestration:
results are further evidence that the reduction of CO2 occurs via physically activated dissolution of serpentine and pH swing process.
HCO3− rather than CO2. Additionally, as shown in Table S2 of Chem. Eng. Sci. 2004, 59 (22−23), 5241−5247.

6008 dx.doi.org/10.1021/es405210d | Environ. Sci. Technol. 2014, 48, 6003−6009


Environmental Science & Technology Article

(3) Meng, M.; Niu, D. Modeling CO2 emissions from fossil fuel (23) Tian, G.; Yuan, H.; Mu, Y.; He, C.; Feng, S. Hydrothermal
combustion using the logistic equation. Energy 2011, 36 (5), 3355− reactions from sodium hydrogen carbonate to phenol. Org. Lett. 2007,
3359. 9 (10), 2019−2021.
(4) Kone, A. C.; Buke, T. Forecasting of CO2 emissions from fuel (24) Cornia, A.; Caneschi, A.; Dapporto, P.; Fabretti, A. C.;
combustion using trend analysis. Renewable Sustainable Energy Rev. Gatteschi, D.; Malavasi, W.; Sangregorio, C.; Sessoli, R. Manganese-
2010, 14 (9), 2906−2915. (III) formate: A three-dimensional framework that traps carbon
(5) CO2 emissions from fuel combustion. Energy Explor. Exploit. dioxide molecules. Angew. Chem., Int. Ed. 1999, 38 (12), 1780−1782.
1997, 15, (6), 523−523. (25) Shabala, S. Metal cations in CO2 assimilation and conversion by
(6) Grimm, N. B.; Chapin, F. S., III; Bierwagen, B.; Gonzalez, P.; plants. JOM 2009, 61 (4), 28−34.
Groffman, P. M.; Luo, Y.; Melton, F.; Nadelhoffer, K.; Pairis, A.; (26) Neelameggham, N. R. Soda-fuel metallurgy: Metal ions for
Raymond, P. A.; Schimel, J.; Williamson, C. E. The impacts of climate carbon neutral CO2 and H2O reduction. JOM 2009, 61 (4), 25−27.
change on ecosystem structure and function. Front. Ecol. Environ. (27) Neelameggham, N. R. Solar pyrometallurgyAn historic
review. JOM 2008, 60 (2), 48−50.
2013, 11 (9), 474−482.
(28) Sheshko, T. F.; Serov, Y. M. Bimetallic systems containing Fe,
(7) Patz, J. A.; Campbell-Lendrum, D.; Holloway, T.; Foley, J. A.
Co, Ni, and Mn nanoparticles as catalysts for the hydrogenation of
Impact of regional climate change on human health. Nature 2005, 438 carbon oxides. Russ. J. Phys. Chem. A 2012, 86 (2), 283−288.
(7066), 310−317. (29) Jin, F. M.; Zhou, Z. Y.; Moriya, T.; Kishida, H.; Higashijima, H.;
(8) Wheeler, T.; von Braun, J. Climate change impacts on global food Enomoto, H. Controlling hydrothermal reaction pathways to improve
security. Science 2013, 341 (6145), 508−513. acetic acid production from carbohydrate biomass. Environ. Sci.
(9) Sumaila, U. R.; Cheung, W. W. L.; Lam, V. W. Y.; Pauly, D.; Technol. 2005, 39 (6), 1893−1902.
Herrick, S. Climate change impacts on the biophysics and economics (30) Jin, F. M.; Kishita, A.; Moriya, T.; Enomoto, H. Kinetics of
of world fisheries. Nat. Clim. Change 2011, 1 (9), 449−456. oxidation of food wastes with H2O2 in supercritical water. J. Supercrit.
(10) Wang, W.; Wang, S.; Ma, X.; Gong, J. Recent advances in Fluids 2001, 19 (3), 251−262.
catalytic hydrogenation of carbon dioxide. Chem. Soc. Rev. 2011, 40 (31) Yasaka, Y.; Yoshida, K.; Wakai, C.; Matubayasi, N.; Nakahara,
(7), 3703−3727. M. Kinetic and equilibrium study on formic acid decomposition in
(11) Yadav, R. K.; Baeg, J.-O.; Oh, G. H.; Park, N.-J.; Kong, K.-j.; relation to the water-gas-shift reaction. J. Phys. Chem. A 2006, 110
Kim, J.; Hwang, D. W.; Biswas, S. K. A photocatalyst−enzyme coupled (38), 11082−11090.
artificial photosynthesis system for solar energy in production of (32) Yu, J. L.; Savage, P. E. Decomposition of formic acid under
formic acid from CO2. J. Am. Chem. Soc. 2012, 134 (28), 11455− hydrothermal conditions. Ind. Eng. Chem. Res. 1998, 37 (1), 2−10.
11461. (33) Jin, F.; Yun, J.; Li, G.; Kishita, A.; Tohji, K.; Enomoto, H.
(12) Jessop, P. G.; Joo, F.; Tai, C. C. Recent advances in the Hydrothermal conversion of carbohydrate biomass into formic acid at
homogeneous hydrogenation of carbon dioxide. Coord. Chem. Rev. mild temperatures. Green Chem. 2008, 10 (6), 612−615.
2004, 248 (21−24), 2425−2442. (34) Takahashi, H.; Liu, L. H.; Yashiro, Y.; Ioku, K.; Bignall, G.;
(13) Roy, S. C.; Varghese, O. K.; Paulose, M.; Grimes, C. A. Toward Yamasaki, N.; Kori, T. CO2 reduction using hydrothermal method for
solar fuels: Photocatalytic conversion of carbon dioxide to hydro- the selective formation of organic compounds. J. Mater. Sci. 2006, 41
carbons. ACS Nano 2010, 4 (3), 1259−1278. (5), 1585−1589.
(14) Varghese, O. K.; Paulose, M.; LaTempa, T. J.; Grimes, C. A.
High-rate solar photocatalytic conversion of CO2 and water vapor to
hydrocarbon fuels. Nano Lett. 2009, 9 (2), 731−737.
(15) Indrakanti, V. P.; Kubicki, J. D.; Schobert, H. H. Photoinduced
activation of CO2 on Ti-based heterogeneous catalysts: Current state,
chemical physics-based insights and outlook. Energy Environ. Sci. 2009,
2 (7), 745−758.
(16) Takeda, H.; Koike, K.; Inoue, H.; Ishitani, O. Development of
an efficient photocatalytic system for CO2 reduction using rhenium(I)
complexes based on mechanistic studies. J. Am. Chem. Soc. 2008, 130
(6), 2023−2031.
(17) Uchida, S. Solar power concentration using laser technology for
the magnesium energy circulation. Proceedings of the 18th SolarPACES
Conference; Marrakech, Morocco, Sept 11−14, 2012; I-11.
(18) Haueter, P.; Moeller, S.; Palumbo, R.; Steinfeld, A. The
production of zinc by thermal dissociation of zinc oxideSolar
chemical reactor design. Sol. Energy 1999, 67 (1−3), 161−167.
(19) Galvez, M. E.; Loutzenhiser, P. G.; Hischier, I.; Steinfeld, A.
CO2 splitting via two-step solar thermochemical cycles with Zn/ZnO
and FeO/Fe3O4 redox reactions: Thermodynamic analysis. Energy
Fuels 2008, 22 (5), 3544−3550.
(20) Hu, Y. H. A highly efficient photocatalyst hydrogenated black
TiO2 for the photocatalytic splitting of water. Angew. Chem., Int. Ed.
2012, 51 (50), 12410−12412.
(21) Wu, B.; Gao, Y.; Jin, F.; Cao, J.; Du, Y.; Zhang, Y. Catalytic
conversion of NaHCO3 into formic acid in mild hydrothermal
conditions for CO2 utilization. Catal. Today 2009, 148 (3−4), 405−
410.
(22) Jin, F.; Gao, Y.; Jin, Y.; Zhang, Y.; Cao, J.; Wei, Z.; Smith, R. L.,
Jr. High-yield reduction of carbon dioxide into formic acid by zero-
valent metal/metal oxide redox cycles. Energy Environ. Sci. 2011, 4 (3),
881−884.

6009 dx.doi.org/10.1021/es405210d | Environ. Sci. Technol. 2014, 48, 6003−6009

You might also like