You are on page 1of 7

Journal of Catalysis 362 (2018) 78–84

Contents lists available at ScienceDirect

Journal of Catalysis
journal homepage: www.elsevier.com/locate/jcat

Priority Communication

Additive free, room temperature direct homogeneous catalytic carbon


dioxide hydrogenation in aqueous solution using an iron(II) phosphine
catalyst
Mickael Montandon-Clerc, Gábor Laurenczy ⇑
Institut des Sciences et Ingénierie Chimiques, École Polytechnique Fédérale de Lausanne (EPFL), CH-1015 Lausanne, Switzerland

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

Article history: The negative consequences of the global warming require an important reduction of CO2 emission; and
Received 31 January 2018 the valorization of the carbon dioxide, its transformation into useful chemicals is essential. We present
Revised 22 March 2018 here our studies on the direct CO2 hydrogenation reaction, yielding formic acid. In water, for the first
Accepted 26 March 2018
time, an Fe(II) catalyst using meta-trisulfonated-tris[2-(diphenyl-phosphino)-ethyl]phosphine (PP3TS)
Available online 16 April 2018
ligand, has been found active in CO2 reduction. In homogeneous catalytic reactions, without any addi-
tives, at room temperature, under hydrogen and carbon dioxide gas pressures up to 0.5 M of formic acid
Keywords:
is obtained, in acidic aqueous solutions. The same catalyst is active also in the reverse reaction, under dif-
Carbon dioxid hydrogenation
Iron(II) catalyst
ferent reaction conditions, i.e. at low pressure and high temperature. The CO2 reduction and formic acid
Direct CO2 reduction dehydrogenation catalytic cycle has been repeated several times; without deactivation of the catalyst, it
Aqueous solution is not sensitive to oxygen/air. The Fe(II)-PP3TS complex could be a suitable catalyst in a chemical hydro-
Formic acid gen storage/delivery system.
Formic acid dehydrogenation Ó 2018 Elsevier Inc. All rights reserved.
Hydrogen storage

1. Introduction organisms. CO2 can be also used as hydrogen vector, to store


chemically the H2 gas, an elegant contribution of CO2 utilization.
Global warming and all the consequences, climate changes are Hydrogen is one of the most promising energy carriers. It has
the reality today. There is a consensus among the majority of sci- high gravimetric energy density and clean combustion pathways,
entists concerning the cause: augmentation of CO2 level in the but it is difficult to store it. Conventional hydrogen storage meth-
atmosphere (over 400 ppm today versus 345 thirty five years ods, like high pressure tanks and liquefaction, have safety and cost
ago) [1]. CO2 is a greenhouse gas, the heat is conserved at the sur- issues [5]. Therefore, the chemical hydrogen storage, the catalytic
face of Earth. However, this is not the only issue, with more CO2 in reduction of CO2 into formic acid, into methanol or into other use-
the atmosphere, the pH of the sea water will decrease, leading to ful chemicals have both the advantage of storing hydrogen and
the acidification of the oceans [2]. All these changes could have adding value to carbon dioxide.
tremendous repercussions on the ecosystem of our planet. In Formic acid (FA) has recently drawn attention in H2 storage
response, both politicians and scientists should make effort to limit because its several advantages: It is liquid at room temperature
the carbon dioxide emission. with a volumetric hydrogen storage capacity of 53 g/L (it has a
The main source of the CO2 emission is the fossil fuel combus- gravimetric hydrogen content of 4.4 wt%). FA has low toxicity
tion. Increasing utilization of the renewable energy sources (hydro- and it is non-flammable below 85 vol% concentration (diluted with
electric, geothermic, wind and solar energy) is the first step toward water), it is a good candidate for hydrogen storage [6,7]. Since
the reduction of CO2 production. An interesting, although minor 2006, a large number of studies dealt with the selective FA dehy-
contribution would be to use CO2 in chemical syntheses. CO2 can drogenation [8,9]. A wide variety of metals and ligands have been
be used/is used as a C1 building block for useful molecules (e.g. used as catalysts for homogeneous and heterogeneous formic acid
urea, salicylic acid, cyclic carbonates, epoxides, formaldehyde, splitting into H2 and CO2, using either noble or non-noble metals,
etc) [3,4]. Nature uses CO2 as carbon source for the living in a wide selection of solvents, reviewed recently [10].
The opposite reaction, the direct carbon dioxide hydrogenation
to FA has got special attention as well, a large number of catalysts
⇑ Corresponding author. have been reported for the reduction of CO2 [10]. However, in the
E-mail address: gabor.laurenczy@epfl.ch (G. Laurenczy). overwhelming majority of the cases, this reduction has been

https://doi.org/10.1016/j.jcat.2018.03.030
0021-9517/Ó 2018 Elsevier Inc. All rights reserved.
M. Montandon-Clerc, G. Laurenczy / Journal of Catalysis 362 (2018) 78–84 79

carried out in the presence of a base and/or use carbonate and the saturation concentrations. FA concentrations were measured
bicarbonate substrates instead of CO2 gas. Indeed, CO2 has a high in situ under pressure by quantitative 1H and 13C NMR (see
thermodynamic stability; in gas phase it cannot be directly hydro- Experimental).
genated to FA (the Gibbs free energy for this reaction, DG is posi- When we have pressurized an aqueous solution of Fe(BF4)2
tive, +33 KJ/mole). In basic aqueous solutions, CO2 is transformed (0.05 M) and PP3TS (0.05 M) with 50 bar of H2 and 50 bar of CO2
into bicarbonates (HCO 2
3 ) and carbonates (CO3 ), and these sub-
in a 10 mm sapphire NMR tube; heated this tube to 50 °C and
strates can be reduced directly, the Gibbs free energy change then mixed it for several hours, more than 0.1 M formic acid solution
will be negative (around – 35 KJ/mole), hence favorable for the was detected by 1H and 13C NMR in the solution.
reaction [11]. We have investigated the effect of the catalyst concentration on
Nevertheless, there are only a few reports for the direct reduc- the formic acid production/yield, and the iron(II) to phosphine
tion of CO2 in acidic solutions, acidic conditions [10]. All these ligand ratio, using Fe(BF4)2 salts as catalyst precursors. Pressure
studies use, however, noble metal based catalysts and to the best (50 bar H2 and 50 bar CO2) and temperature (25 °C) were kept con-
of our knowledge there is no report for direct, homogeneous CO2 stant during the whole reaction (generally 60 h – checking several
hydrogenation, using a first row transition metal (non-noble times the formic acid concentration, verifying to arrive to equilib-
metal) catalyst, in water. Combining both hydrogenation and ria), Table 1. The ratio of the iron salt with respect to the ligand was
dehydrogenation reactions, applying a cheap iron(II) complex kept constant while varying the concentration of the catalyst
could lead to the development of an affordable, battery-like complex.
energy/hydrogen storage system. This system could have a poten- The catalyst concentration does not influence the formic acid
tial application in mobile energy supply. yield significantly (Table 1), indicating that the limiting factor for
the conversion is thermodynamic. It should be the hydrogen and
CO2 equilibrium concentration under pressure in water. The cata-
2. Results lyst concentration certainly influences the reaction rate.
The reaction requires gas solubilization under pressure (CO2
Recently we have studied the homogeneous formic acid dehy- and H2), it is necessity to have an intensive mixing. The quantita-
drogenation reaction with the Fe(II)-PP3TS catalyst [12]. Now we tive NMR technique gave the only possibility for FA concentration
have investigated the direct reduction of CO2 in aqueous solution determination in situ, under pressure (time-consuming measure-
(that is in acidic medium), using the same iron(II) complex. ments, long acquisitions). For these reasons we were not able to
When we have pressurized with hydrogen and carbon dioxide do precise kinetic analysis for this reaction. We are working on
the aqueous solution of Fe(BF4)2; in presence of meta-trisulfona to find a suitable experimental setup to collect detailed kinetic
ted-tris[2-(diphenylphosphino)ethyl]phosphine sodium salt (PP3- information on this reaction.
TS – Fig. 1); we could detect the formation of formic acid in the In Table 2 the results concerning the effect of the ligand to iron
reaction mixture, even at room temperature. 10 mm high pressure (II) ratio has been shown. The ratio of the iron salt with respect to
sapphire NMR tubes were used to carry out these reactions, a mix- the ligand was varied between 2:1 and 1:3. Different iron salt were
ing device was applied to solubilize H2 and CO2, to reach always tested as precursors (FeCl2, Fe(BF4)2 and (NH4)2Fe(SO4)2. A poison-
ing effect has occurred when FeCl2 was used (at high chloride con-
centrations the formation of chloride bridged dimers is possible),
the catalytic activity of this precursor were lower.
Using half equivalent of PP3TS ligand could lead to bimetallic
complex formation, with reduced catalytic activity. In case of
large PP3TS ligand excess, the formation of the ML2, ML3 type
complexes could reduce the available active coordination sites
around the Fe(II) ion. The excess of the PP3TS ligand blocks the
access of the substrate to the iron(II) cation, avoiding catalytic
reduction.
As it is can be seen from variation of the chemical shift of the 31P
signal of the tris-ethylene linked P atom in meta-trisulfonated-tris
[2-(diphenylphosphino)ethyl]phosphine (PP3TS) at acidic and neu-
tral solutions, this P atom can be protonated (Fig. 2a), while the
other PPh2 type P atoms not, in the studied pH range. Indeed, the
31
P NMR peak, corresponding to the triethylenephosphine bridge-
head phosphorus, it was shifted from 20 ppm to 15 ppm under
acidic conditions with decreasing pH, while the chemical shift of
Fig. 1. Meta-trisulfonated-tris[2-(diphenylphosphino)ethyl]phosphine sodium salt the signal corresponding to PPh2 phosphorus atoms practically
(PP3TS, 1). does not change. From the 31P NMR shifts it was possible to

Table 1
Formic acid yield dependence on the catalyst concentration.

Fe(BF4)2 concentration [mM] 8.3 10 15 25 30 50 75 100


Final formic acid conc. [M] 0.21 0.13 0.19 0.27 0.21 0.30 0.23 0.25
0.25 0.18 0.20 0.27 0.24 0.22 0.22 0.20
0.20 0.15 0.23 0.28 0.21 0.23 0.23 0.23
Average 0.22 0.15 0.21 0.28 0.22 0.25 0.23 0.23
stda 0.04 0.04 0.04 0.03 0.03 0.06 0.03 0.04

PP3TS/Fe(II) ratio = 1; 50 bar H2 and 50 bar CO2, room temperature, final pH 2.0–2.5, reaction time: 60 h.
a
Estimated standard deviation, reproducibility.
80 M. Montandon-Clerc, G. Laurenczy / Journal of Catalysis 362 (2018) 78–84

Table 2
Effect of PP3TS to iron(II) ratio on the final formic acid concentration.

Ligand Ratio 0.5 1 2 3


Formic acid production [M] 0.067 0.27 0.072 0.082 Fe(BF4)2
0.072 0.27 0.07 0.092
0.071 0.28 0.07 0.087
0.035 0.055 0.22 0.041 (NH4)2Fe(SO4)2
0.030 0.053 0.22 0.045
0.031 0.054 0.22 0.046
– 0.010 0.0078 – FeCl2
– 0.0094 0.011 –
– 0.0082 0.0091 –
Average 0.051 0.16 0.14 0.066
stda 0.019 0.11 0.09 0.03

Iron(II) salt concentration: 25 mM, 50 bar H2 and 50 bar CO2, room temperature, final pH 2.0–2.5, average of 6 to 10 measurements, reaction time: 60 h.
a
Estimated standard deviation, reproducibility.

Fig. 2. (a) 161 MHz 31P NMR spectra of the PP3TS ligand at pH = 1 and pH = 7 (b) Chemical shifts of the 31P signal of the single tri-ethylene linked P in the meta-trisulfonated-
tris[2-(diphenylphosphino)ethyl]phosphine, PP3TS as a function of pH) atom.
M. Montandon-Clerc, G. Laurenczy / Journal of Catalysis 362 (2018) 78–84 81

Table 3
Formic acid production as a function of H2 and CO2 gas pressure and H2/CO2 pressure ratio.

H2/CO2 [Bar] 25/25 50/25 50/50 100/50


Formic acid production [M] 0.028 0.066 0.27 0.50
0.022 0.080 0.27 0.40
– 0.069 0.28 0.47
Average 0.025 0.072 0.28 0.46
stda 0.01 0.02 0.02 0.06

Fe(BF4)2 concentration: 25 mM, PP3TS ratio: 1, room temperature, final pH 2.0–2.5, reaction time: 60 h.
a
Estimated standard deviation, reproducibility.

From the data collected in Table 4, the enthalpy of the carbon


Table 4
Formic acid yield as a function of the reaction temperature.
dioxide hydrogenation reaction can be determined (Fig. 3).
 
If we consider that Dr G ¼ RTlnðKÞ and that Dr G ¼ DH  T DS,
Temperature [°C] 25 55 65 85
with K ¼ ½H2½FA 
½CO2 
, one can plot lnð½FAÞ v s 1=T, where Dr H
R
will be the
Formic acid production [M] 0.27 0.13 0.075 0.054
0.27 0.11 0.069 0.041 slope. The slope being 3163.1; one can find Dr H ¼
0.28 0.13 0.071 0.048 26:29kJ=K  mol for this reaction under our conditions.
Average 0.28 0.12 0.072 0.048 Thermodynamic considerations: Considering that
stda 0.03 0.02 0.006 0.008 ½HCOOH
the Gibbs free energy; DrG ¼ RTlnðK therm Þ, that K therm ¼ ½H 2 ½CO2 
,
Fe(BF4)2 concentration: 25 mM, PP3TS ratio: 1, 50 bar H2 and 50 bar CO2, final pH 2– DG
 
2.5, reaction time: 60 h.
rearranging for ½HCOOH ¼ ½H2 ½CO2  exp  RT . With DrG ¼ 4 mol KJ

a
Estimated standard deviation, reproducibility. for formic acid, R ¼ 0:008314 Kmol
LBar
, T ¼ 298K at room temperature,
½H2  ¼ 0:08 M (estimated using Henry’s law with Kh = 1279.31
Latm/mol [13] and ½CO2  ¼ 1:25 M [14] at 50 bar, one can calculate
calculate the pKa of the tris-ethylene-P phosphine, being pKa = 4.73 a maximum formic acid yield around of 0.5 M under these specific
(Fig. 2b) conditions.
The effect of CO2 and H2 pressure on the formic acid yield, as As it has been shown previously, (1) can catalyze efficiently the
well as the consequence of the changing the proportion of CO2 to formic acid dehydrogenation [15]. This catalyst stays active after
H2 pressures have been also investigated (Table 3). The highest several cycles of hydrogenation/dehydrogenation (Table 5).
FA concentration was obtained using 100 bar of H2 and 50 bar of
CO2. Higher concentrations of formic acid were obtained using
higher hydrogen and carbon dioxide pressures (higher dissolved Table 5
hydrogen and CO2 concentrations are increasing the FA yield – Le Recycling experiments, reuse of catalyst 1 (for details see the Experimental section).
Chatelier-Braun principle). The measurements under 150 bars total Entry Cycle formic acid concentration [M] pHfinal
pressure were carried out in an autoclave (these sapphire NMR
1 1 0.22 2.0
tubes can be used in up to 100 bar total pressure). 2 2 0.18 2.0
The direct carbon dioxide hydrogenation is an exothermic reac- 3 3 0.20 2.5
tion, the temperature influences the produced formic acid quan-
Fe(BF4)2: 25 mM, PP3TS to iron ratio is 1:1, using 50 bar H2 and 50 bar CO2, room
tity, concentration. As this reduction is exothermic, increasing temperature, initial pH = 3.5, average of 3 to 6 measurements, reproducibility ±
temperature decreases the FA yield (Table 4). 10%.

-0.5

-1

-1.5
ln([FA])

-2

-2.5

-3

-3.5
0.00275 0.00285 0.00295 0.00305 0.00315 0.00325 0.00335 0.00345
1/T
Fig. 3. Plot of the natural logarithm of the formic acid concentration, yielded in the direct hydrogenation reaction, as a function of the reciprocal reaction temperature, 1/T.
The linear fit resulted in a slope of 3163.1 which can then be used to calculated the enthalpy of reaction.
82 M. Montandon-Clerc, G. Laurenczy / Journal of Catalysis 362 (2018) 78–84

2.1. Reaction mechanism sulfonated tripodal ligand, P(CH2CH2PPh2)3, PP3, has been studied
in details by Bianchini and coworkers in acetone-d6, allowing them
We have tried to get mechanistic information on the direct car- to take the 1H NMR spectra down to – 85 °C [37,38]. At the temper-
bon dioxide hydrogenation reaction, using the Fe(II)-PP3TS com- ature range of 10 °C K to +45 °C their 1H NMR spectra are very
plex as catalyst. We have pressurized a 0.025 M catalyst (1) similar to our observations, showing the typical intramolecular
solution (in H2O and also in D2O) using medium pressure sapphire exchange between of the coordinated dihydrogen and hydride. At
NMR tubes, up to 100 bar gas pressures (mixtures of CO2 and H2). lower temperatures they could see a well resolved multiplet
Unfortunately, in the accessible temperature and pressure range in (tdd) at - 12.47 ppm (intensity: 1 H), which was assigned to a ter-
water (between 1 °C and 110 °C; and up to 100 bar – 50 bar of CO2 minal hydride which is coupled with the phosphorus nuclei of PP3.
and 50 bar of H2 pressures), we have not get any signal, which At 85 °C the resonance of the coordinated dihydrogen (intensity:
could help to identify catalytically active species or intermediates 2 H) ligand started to broaden, but not resolved to give the possi-
in the 1H and 13C NMR spectra. In the 1H NMR spectra only the dis- bility to determine the coupling constants. They have also detected
solved H2, the ligand protons, the formed formic acid and the water the hydride – dihydrogen exchange. The same hydrido-dihydrogen
signal could be seen. In the 13C NMR, after taking a large number of complex has been identified by Beller and coworkers in the cat-
scans, only the dissolved CO2, the resulting FA and the carbon alytic cycle during the formic acid dehydrogenation [14]. A very
atoms of the ligand could be detected. similar behavior was already described by Morris et al. for trans-
To identify the possible catalytically active species, hydrides, we [Fe(g2-H2)(H)(dppe)2]BF4 (dppe = PPh2CH2CH2PPh2) and for
have repeated the experiments with the catalyst (1) and only with trans-[Ru(g2-H2)(H)(dppe)2]BF4 [15], giving further confirmation
pure H2 pressure. Pressurizing the Fe(II)-PP3TS catalyst solution to our proposition. In order to determine the presence of a non-
(0.025 M) at 100 bar of H2, the solution shifted from purple to dark classical dihydrogen ligand, as well as to confirm the fast exchange
pink and one broad signal in the hydride region could be seen at between the hydride and the dihydrogen ligands, the respective T1
around 10 ppm at 80 °C (Fig. 4). Cooling down gradually the sam- relaxation times were measured. The T1 value for the signal at
ple lead to the formation of two broad signals, at around 8.2 and 8.2 ppm (supposed to be a non-classical dihydrogen ligand)
13.5 ppm, the chemical shifts and integrals suggesting the ratio was found to be 30 ms while the relaxation time, T1 for the signal
of the corresponding hydrogens of 2 to 1. Despite of our tentative at 13.5 ppm (supposed to be a hydride ligand) was found to be
and our efforts, these signals stayed broad and we could not get 50 ms. Usually the T1 value for this latter one is somewhat longer,
any further information for the multiplicities and couplings. Fortu- the determined value certainly short because of the intramolecular
nately, a very similar system, the iron(II) with the same, but not exchange [16,17].

Fig. 4. 400 MHz 1H NMR spectra of a 0.025 M Fe(II)-PP3TS solution (hydride region) under 100 bar of H2 as a function of temperature (for the first spectrum t = 80 °C, and the
temperature is decreasing with 10 °C for the following ones).
M. Montandon-Clerc, G. Laurenczy / Journal of Catalysis 362 (2018) 78–84 83

3. Conclusion expected C = 51.64% and H = 4.02% for C42H39Na3O9P4S3; obtained


C = 51.64% and H = 4.01%.
We have shown that the direct homogeneous catalytic CO2
hydrogenation in aqueous solution, at room temperature, without 5. Isochoric experiments
any additives can be realized using the PP3TS ligand and iron(II) salts
as catalyst precursors. The catalyst is robust and can be used for In a typical experiment, PP3TS (0.0245 g, 0.025 mmol) was dis-
multiple cycles, its activity does not decrease after contact with solved in water (2.0 mL) in a 10 mm NMR sapphire tube, and one
air/oxygen. It has been shown that the temperature and the pressure equivalent of Fe(II) salt was added and dissolved (FeCl2, Fe(BF4)2
are the key parameters and influence the synthetized final formic or (NH4)2Fe(SO4)2). The tube was then sealed and pressurized with
acid yield/concentration. This hydrogenation reaction is reversible CO2 and H2 (50–50 bar). The samples were continuously shaken
in presence of the iron(II) – PP3TS catalyst, the equilibrium position, with a rate of 200/min. This ensures the H2 and CO2 gas dissolution
the direction of the reaction depends on the temperature and the H2 and saturation. Reactions were followed using quantitative 1H and
and CO2 pressures. Using this system, a charge/discharge hydrogen 13
C NMR by monitoring the formic acid peaks, relative to DSS stan-
battery, a chemical hydrogen cylinder can be foreseen. The enthalpy dard. NMR spectra were taken upon defined interval (approxi-
of the formic acid formation and the activation energy for the carbon mately every 4 h). When the integral of the formic acid peaks
dioxide hydrogenation have been determined. relative to the DSS remained identical (within 5% difference) after
three subsequent measurements, equilibrium was believed to be
reached, usually after 60 h at room temperature.
4. Experimental

4.1. General 6. Recycling experiments

The air-sensitive compound (PP3) was manipulated under nitro- A 10 mm sapphire NMR tube with the catalyst (0.025 M) was
gen atmosphere using standard Schlenk techniques. PP3 (97%) was prepared as previously described, pressurized (50 bar CO2 and
purchased from Acros. All other reagents were commercially avail- 50 bar H2), and shaked. Once equilibrium was reached, the pres-
able and used as received. sure was release and the residual H2 and CO2 gases were vented.
The open tube was thermostatted at 80 °C until no more formic
acid could be detected in the solution by quantitative 1H and 13C
4.2. NMR spectroscopy NMR. The tube was then cooled back to room temperature and this
1
process was repeated.
H, 13C and 31P NMR spectra were recorded on Bruker DRX-400
(5 mm) and Avance DRX-400 (10 mm) instruments. 3-(Trimethylsi
Acknowledgements
lyl)-1-propanesulfonic acid sodium salt (DSS) was used as an inter-
nal standard for 1H and 13C NMR experiments. For quantitative
The Swiss Competence Center for Energy Research (SCCER), the
NMR measurements long acquisition delay times (D1 > 6 * T1) were
Swiss Commission for Technology and Innovation (CTI) – Switzer-
used [18]. The spectra were processed with TopSpin, MestReNova
land and the École Polytechnique Fédérale de Lausanne (EPFL) are
and Igor Pro softwares.
thanked for financial support.

4.3. PP3TS synthesis [15] References

The synthesis of the ligand was performed as described previ- [1] CO2 now, (n.d.). http://co2now.org/.
[2] S. Cooley, H. Kite-Powell, S. Doney, Ocean acidification’s potential to alter
ously. A 3 L three-neck flask was equipped with a magnetic stirrer global marine ecosystem services, Oceanography 22 (2009) 172–181, https://
and charged with H2SO4 (98%, 56 mL, 1.05 mol), cooled to 10 °C doi.org/10.5670/oceanog.2009.106.
by an ice-salt bath. PP3 (5g, 7.45 mmol) was slowly added over a [3] Q. Liu, L. Wu, R. Jackstell, M. Beller, Using carbon dioxide as a building block in
organic synthesis, Nat. Commun. 6 (2015) 5933, https://doi.org/10.1038/
1 h period. After complete dissolution of the solid phosphine,
ncomms6933.
oleum (65% SO3, 30 mL, 0.33 mol) was carefully added to the reac- [4] T.G. Ostapowicz, M. Schmitz, M. Krystof, J. Klankermayer, W. Leitner, Carbon
tion mixture. The reaction mixture was allowed to warm to room dioxide as a C1 building block for the formation of carboxylic acids by formal
temperature and stirred for 4 days. After this time the reaction catalytic hydrocarboxylation, Angew. Chemie Int. Ed. 52 (2013) 12119–12123,
https://doi.org/10.1002/anie.201304529.
was quenched with ice (100 g) and the mixture was neutralized [5] A. Léon, Hydrogen Technology in Mobile and Portable Applications, Springer-
over a 4 h period by adding NaOH (7M) with a dropping funnel. Verlag, Berlin, Heidelberg, 2008.
The water was removed under vacuum and MeOH (altogether 1 [6] S. Enthaler, J. von Langermann, T. Schmidt, Carbon dioxide and formic acid—
the couple for environmental-friendly hydrogen storage?, Energy Environ Sci.
L) added to extract the sulfonated phosphine salt. The mixture 3 (2010) 1207, https://doi.org/10.1039/b907569k.
was filtered under nitrogen and the filtrate reduced to dry, to [7] F. Joó, Breakthroughs in hydrogen storage-formic acid as a sustainable storage
afford 4.2 g (58%) of pure PP3TS as purple crystals. material for hydrogen, ChemSusChem. 1 (2008) 805–808, https://doi.org/
10.1002/cssc.200800133.
The 31P NMR spectrum of PP3TS displays one broad signal from [8] C. Fellay, P.J. Dyson, G. Laurenczy, Hydrogen production from formic acid, EP
22.07 ppm to 18.26 ppm for the bridgehead phosphorous, while 1918247, 2006.
the diphenylphosphino groups gave rise to a broad signal from [9] C. Fellay, P.J. Dyson, G. Laurenczy, A viable hydrogen-storage system based on
selective formic acid decomposition with a ruthenium catalyst, Angew.
15.52 to 11.54 ppm. Phosphine oxides were not observed dur- Chemie - Int. Ed. 47 (2008) 3966–3968, https://doi.org/10.1002/
ing the synthesis in the 31P NMR spectra (30–80 ppm). The 1H anie.200800320.
NMR spectrum contains a broad resonance centred at 1.43 ppm [10] K. Sordakis, C. Tang, L.K. Vogt, H. Junge, P.J. Dyson, M. Beller, G. Laurenczy,
Homogeneous catalysis for sustainable hydrogen storage in formic acid and
for the three equivalent ethylene groups and several broad peaks
alcohols, Chem. Rev. 118 (2018) 372–433, https://doi.org/10.1021/acs.
between 6.22 and 8.44 ppm for the phenyl rings. The respective chemrev.7b00182.
integral ratio was very near to 12:27, indicating that three phenyl [11] G. Laurenczy, F. Joó, L. Nádasdi, Formation and characterization of water-
groups were sulfonated. The negative mode ESI (electrospray ion- soluble hydrido-Ruthenium(II) complexes of 1,3,5-Triaza-7-
phosphaadamantane and their catalytic activity in hydrogenation of CO2 and
ization) MS corresponds to a trisulfonated salt (peaks at m/z = HCO3 - in aqueous solution, Inorg. Chem. 39 (2000) 5083–5088, https://doi.
302.98 for [M]3 and 465.72 for [M+Na]2). Elemental analysis: org/10.1021/ic000200b.
84 M. Montandon-Clerc, G. Laurenczy / Journal of Catalysis 362 (2018) 78–84

[12] M. Montandon-Clerc, A.F. Dalebrook, G. Laurenczy, Quantitative aqueous [16] C. Bianchini, M. Peruzzini, F. Zanobini, An exceptionally stable cis-(hydride)
phase formic acid dehydrogenation using iron(II) based catalysts, J. Catal. 343 (g2-dihydrogen) complex of iron, J. Organomet. Chem. 354 (1988) C19–C22,
(2016) 62–67, https://doi.org/10.1016/j.jcat.2015.11.012. https://doi.org/10.1016/0022-328X(88)87057-8.
[13] D.W. Green, R.H. Perry, Solubilities, Perry’s Chem. Eng. Handbook, Eighth Ed. [17] C. Bianchini, F. Laschi, M. Peruzzini, F.M. Ottaviani, A. Vacca, P. Zanello,
(2008). Electrochemistry as a diagnostic tool to discriminate between classical M-(H)2
[14] Z. Duan, R. Sun, C. Zhu, I.M. Chou, An improved model for the calculation of and nonclassical M-(H2) structures within a family of dihydride and
CO2 solubility in aqueous solutions containing Na+, K+, Ca2+, Mg2+, Cl, and SO24, dihydrogen metal complexes, Inorg. Chem. 29 (1990) 3394–3402, https://
Mar. Chem. 98 (2006) 131–139, https://doi.org/10.1016/ doi.org/10.1021/ic00343a025.
j.marchem.2005.09.001. [18] S. Moret, P.J. Dyson, G. Laurenczy, Direct, in situ determination of pH and
[15] R.H. Morris, J.F. Sawyer, M. Shiralian, J. Zubkowski, Two molecular hydrogen solute concentrations in formic acid dehydrogenation and CO(2)
complexes: trans-[M(.eta.2-H2)(H)(PPh2CH2CH2PPh2)2]BF4 (M = Fe, Ru). The hydrogenation in pressurised aqueous solutions using (1)H and (13)C NMR
crystal structure determination of the iron complex, J. Am. Chem. Soc. 107 spectroscopy, Dalton Trans. 42 (2013) 4353–4356, https://doi.org/10.1039/
(1985) 5581–5582, https://doi.org/10.1021/ja00305a071. c3dt00081h.

You might also like