You are on page 1of 4

Communication

Cite This: J. Am. Chem. Soc. XXXX, XXX, XXX−XXX pubs.acs.org/JACS

Integrative CO2 Capture and Hydrogenation to Methanol with


Reusable Catalyst and Amine: Toward a Carbon Neutral Methanol
Economy
Sayan Kar, Raktim Sen, Alain Goeppert, and G. K. Surya Prakash*
Loker Hydrocarbon Research Institute and Department of Chemistry, University of Southern California, University Park, Los
Angeles, California 90089-1661, United States
*
S Supporting Information

condition in the presence of dimethylamine (Figure 1).7 The


ABSTRACT: Herein we report an efficient and recyclable presence of an amine provides the opportunity to capture and
system for tandem CO2 capture and hydrogenation to
methanol. After capture in an aqueous amine solution,
CO2 is hydrogenated in high yield to CH3OH (>90%) in a
biphasic 2-MTHF/water system, which also allows for easy
separation and recycling of the amine and catalyst for
multiple reaction cycles. Between cycles, the produced
methanol can be conveniently removed in vacuo. Employ- Figure 1. Amine assisted CO2 hydrogenation to CH3OH.
ing this strategy, catalyst Ru-MACHO-BH and polyamine
PEHA were recycled three times with 87% of the methanol
producibility of the first cycle retained, along with 95% of
hydrogenate CO2 in tandem, as was demonstrated by our
catalyst activity after four cycles. CO2 from dilute sources
group.8 Since Sanford et al.’s initial report, multiple studies have
such as air can also be converted to CH3OH using this
been published using various metal complexes for amine
route. We postulate that the CO2 capture and hydro-
assisted hydrogenation of CO2 to CH3OH.8,9 However, the
genation to methanol system presented here could be an
integration of CO2 capture with subsequent hydrogenation to
important step toward the implementation of the carbon
CH3OH with easy recycling of the active elements had not
neutral methanol economy concept.
been explored.
Most of the reported metal complexes catalyzing CO2
hydrogenation to CH3OH are soluble in organic solvents,
T he rise of atmospheric CO2 concentration and associated
global warming have prompted researchers to develop
strategies for capturing CO2 from both emission point sources
whereas the capturing amines are soluble in water. Aqueous
solutions of amines have long been utilized for scrubbing CO2
from industrial gas streams.10 Water is a desirable solvent due
and diffuse sources like ambient air.1 Whereas the captured to its benign nature and ability to enhance the amines’ CO2
CO2 can be sequestered underground in geological formations, absorption capacity. Thus, a biphasic system was envisioned,
a more sustainable approach is to utilize the CO2 to produce where after the hydrogenation step, the amine and catalyst can
fuels and other value-added products.2 CH3OH in particular be easily separated and recycled from the aqueous and organic
can be used as a fuel, fuel additive or precursor in organic layer, respectively (Figure 2). The formed CH3OH can be
synthesis.3 The utilization of CO2 to produce CH3OH through extracted through distillation. Similar biphasic systems were
hydrogenation, followed by the use of CH3OH as fuel results in
an overall carbon neutral cycle, and represents an area of
interest in the context of carbon footprint reduction.4
Development of integrated CO2 capture and utilization
(CCU) systems, wherein the captured CO2 can be directly
converted to value-added products (in this case CH3OH), is an
area of enormous interest as it can bypass the otherwise
intermediary and energy intensive desorption and compression
steps to produce pure CO2. For practical implementation,
recycling of the catalyst and capture material is essential to keep
the entire process cost-effective. Traditional catalysts for CO2
hydrogenation to CH3OH are heterogeneous and require high
temperatures and pressures.5 Over the past decade, however, Figure 2. Schematic representation of biphasic CO2 to methanol
system with recyclable catalyst and amine.
significant advances were made in both indirect and direct one-
pot homogeneous catalytic CO2 to CH3OH synthesis under
much milder conditions.6 In 2015, Sanford et al. demonstrated Received: November 17, 2017
a direct CO2 hydrogenation system to CH3OH under basic

© XXXX American Chemical Society A DOI: 10.1021/jacs.7b12183


J. Am. Chem. Soc. XXXX, XXX, XXX−XXX
Journal of the American Chemical Society Communication

demonstrated recently by us and Leitner et al. for integrated Following the capture, the formed aqueous solutions were
CO2 capture and conversion to formate salts.11 hydrogenated at 145 °C in the presence of homogeneous
For the CO2 capture, amines with low vapor pressures are catalysts and 70 bar of H2, after addition of 5 mL 2-MTHF
desirable to avoid atmospheric amine contamination. High (Table 2). When the catalyst Ru-MACHO-BH (C-1) (10
boiling polyamines were therefore selected for capture, along μmol) was used along with PEHA as the capture material, 47%
with two ethanolamines (Table 1). Among various polyamines, CH3OH yield was observed (5.2 mmol) after 72 h (entry 1).
No concomitant CO/CH4 formation was observed through GC
Table 1. CO2 Capture by Aqueous Amine Solutionsa analysis of the reaction gas mixture. 1H and 13C NMR revealed
that 14% of formed CH3OH was present in the upper organic
layer, whereas the remaining CH3OH along with PEHA,
formamide and formate intermediates, was in the bottom
aqueous layer (Figure S6 and S7). The catalyst remained in the
organic layer as observed by 31P NMR (Figure S5), indicating
the possibility of easy catalyst separation from the biphasic
mixture. Increasing the catalyst loading to 20 μmol increased
the methanol yield to 79% (8.7 mmol) (entry 2). Next, various
amine solutions after CO2 capture (from Table 1) were
hydrogenated to identify the most promising amine for an
integrated CO2 capture/hydrogenation system. Switching from
PEHA to BPEI800 or BPEI25k decreased both the amounts of
CH3OH formed (4.5 and 5.2 mmol, respectively) and the
Entry Amine $/kgb CO2 (mmol)/gc CO2/Nd hydrogenation yield (45% and 50%) (entry 3 and 4). For
1 PEHA 105 (S) 11.0 0.43 LPEI2.5k, LPEI100k, and PAA10k, CH3OH yields decreased
2 BPEI800 352 (S) 10.2 0.46 drastically to 0.9, 0.9, and 0.1 mmol, respectively (entry 5−
3 BPEI25k 320 (S) 10.4 0.47 7). Surprisingly, with MEA, no CH3OH was formed (entry 8),
4e PAA10k 7533 (P) 6.2 0.36 but increased amounts of formamide and formate intermediates
5f LPEI2.5k 56,000 (P) 5.5 0.25 were observed. We surmise that in the presence of primary
6f LPEI100k 24,500 (P) 6.1 0.28 amines, such as PAA10k and MEA, the second hydrogenation
7 MEA 35 (S) 11.7 0.71 step of formamide to methanol becomes more challenging.
8 DEEDA 820 (T) 7.2 0.53 Indeed, when DEEDA, a secondary analogue of MEA, was
a
Capture conditions: Amine (1g), water (3 mL), stirring (800 rpm), used, methanol was obtained in 46% yield after 72 h (3.3
rt. Aqueous amine solutions stirred in CO2 atmosphere at a constant mmol; entry 9). Thus, among various amines, PEHA was the
pressure of 1 psi. Captured CO2 amounts calculated through most efficient for the overall CO2 capture and conversion to
gravimetric analysis. Calculations error ± 5%. bPrices from Sigma- CH3OH. The low vapor pressure and easy availability of
Aldrich (S), Polysciences (P) or TCI America (T), as of Nov 14, 2017. inexpensive PEHA (Table 1) make it promising for a large scale
c
CO2 captured per gram of amine. dmols of CO2 captured per mol of
CCU process.
nitrogen. eCommercial 15 wt % PAA10k aqueous solution used directly.
f
10 mL water, capture at 70 °C Next, known hydrogenation catalysts were screened to
investigate their efficacy. Ru-MACHO (C-2), expectedly, was
almost equally effective to Ru-MACHO-BH, in the presence of
an additional base K3PO4 (entry 10). The P-substituent in the
pentaethylenehexamine (PEHA), and branched polyethyleni- PNP ligand heavily influenced the CH3OH yield. With
mines (BPEI) were found efficient for CO2 capture. The RuHClPNPiPr(CO) (C-3), a meager 5% methanol yield was
aqueous PEHA solution captured 11.0 mmol of CO2 per g of observed (entry 11). Complex MnBrPNPiPr(CO)2 (C-4),
PEHA after 4 h, corresponding to 0.43 mol of CO2 captured recently reported by us to catalyze sequential CO2 hydro-
per mol of amino group (CO2/N), (Table 1, entry 1). The 13C genation to methanol, was only capable of producing CH3OH
NMR of the CO2 loaded aqueous PEHA solution revealed the in 5% yield (0.5 mmol; entry 12). No methanol formed with
presence of carbamate and carbonate/bicarbonate (Figure S2). FeHBrPNPiPr(CO) (C-5) (entry 13). Accumulation of
Similarly, BPEI800 and BPEI25k captured 0.46 and 0.47 CO2/N, formamide and formate intermediates in the case of C-3 to
respectively (10.2 and 10.4 mmol of CO2/g, respectively) C-5 suggests lower activities of these catalysts for the effective
(entry 2−3). CO2 capture by aqueous PAA10k solution was hydrogenation of formamides and formates to CH3OH under
slower, as after 4 h only 6.2 mmol/g of CO2 was captured, the present conditions.
corresponding to 0.36 CO2/N (entry 4). Linear polyethyleni- The most suitable organic solvent for the biphasic hydro-
mines (LPEI2.5k, LPEI100k) displayed limited solubility in water genation system was subsequently explored. A slight decrease in
at room temperature, making them less convenient for aqueous methanol formation was observed when switching from 2-
CO2 capture. With more water (10 mL), and a higher MTHF to cyclopentyl methyl ether (CPME) or p-xylene (52%
temperature (70 °C), LPEI2.5k and LPEI100k captured 5.5 and and 57%, respectively) (entry 14 and 15). The reason behind
6.1 mmol of CO2/g, respectively. Monoethanolamine (MEA), the decrease in methanol yield with more hydrophobic solvents
which has long been used industrially for scrubbing CO2 and is not entirely clear, but is most probably a combination of
H2S from flue gases, was the most effective for CO2 capture different catalyst/H2/CO2/CH3OH solubility in different
both by mass and efficiency of amine utilization (11.7 mmol/g; organic solvents (see SI). Also, the higher hydrophobicity of
0.71 CO2/N) (entry 7). Similarly, 7.2 mmol/g CO2 (0.53 these solvents compared to 2-MTHF resulted in an
CO2/N) was captured by diethanolethylenediamine (DEEDA) accumulation of the produced methanol exclusively in the
(entry 8). aqueous layer (Figure S13). However, the lower solubility of C-
B DOI: 10.1021/jacs.7b12183
J. Am. Chem. Soc. XXXX, XXX, XXX−XXX
Journal of the American Chemical Society Communication

Table 2. Tandem Homogeneous Hydrogenation of CO2 Captured by Aqueous Amine Solutionsa

Entry Amine Captured CO2 (mmol) Catalyst (μmol) Formate (%)b Formamide (%)b MeOH (mmol)b Yield (%)b PMeOHc TON
1 PEHA 11.0 C-1(10) 11 10 5.2 47 0.16 520
2 PEHA 11.0 C-1(20) 5 2 8.7 79 0.17 435
3 BPEI800 10.2 C-1(20) 16 13 4.5 45 0.18 225
4 BPEI25k 10.4 C-1(20) 10 7 5.2 50 0.13 260
5 LPEI2.5k 5.5 C-1(20) 30 15 0.9 16 0.11 45
6 LPEI100k 6.1 C-1(20) 21 44 0.9 15 0.11 45
7 PAA10k 6.2 C-1(20) 32 38 0.1 2 0 5
8 MEA 11.7 C-1(20) 26 17 0 0 nd 0
9 DEEDA 7.2 C-1(20) 15 6 3.3 46 0.14 165
10d PEHA 11.0 C-2(20) 6 13 7.4 67 0.15 370
11d PEHA 11.0 C-3(20) 15 20 0.5 5 0 25
12d PEHA 11.0 C-4(20) 18 19 0.5 5 0 25
13d PEHA 11.0 C-5(20) 20 18 0.0 0 nd 0
14e PEHA 11.0 C-1(20) 12 11 5.7 52 0 285
15f PEHA 11.0 C-1(20) 9 6 6.3 57 0 315
16g PEHA 11.0 C-1(50) 3 0 10.4 95 0.17 208
17h PEHA 5.4 C-1(50) 5 0 4.8 89 0.11 96
a
Reaction conditions: Solutions from Table 1 (as specified) were hydrogenated after adding organic solvent and catalyst. 2-MTHF (5 mL), H2 (70
bar), 145 °C, 72 h. bYields based on 1H NMR with 1,3,5-trimethoxybenzene (TMB) and imidazole (Im) as internal standards for organic and
aqueous layer, respectively. cPMeOH = methanol in organic layer/methanol in aqueous layer. dK3PO4 (1 mmol) added. eCPME used as organic
solvent. fP-xylene used as organic solvent. gH2 (80 bar). hCO2 captured from simulated air (CO2 concentration: 408 ppm) with 0.79 g PEHA. Yield
calculations error ± 5%. TON = mols of methanol formed per mol of catalyst. nd = nondefinable

1 in these solvents at room temperature caused some of the


catalyst to precipitate out from the solution during workup,
making its complete recycling challenging. Hence, 2-MTHF
was identified as the most convenient solvent for repeated
capture and utilization studies. Using 2-MTHF, a CH3OH yield
as high as 95% was obtained with a higher C-1 loading of 50
μmol (entry 16). Finally, CO2 from air was also captured and
hydrogenated to CH3OH in high yields following this protocol
(entry 17).
With the optimized selection of amine (PEHA), catalyst (C- Figure 3. Methanol formation with catalyst recycling (A) and catalyst
1), and organic solvent (2-MTHF), two recycling studies were and amine recycling (B). Reaction conditions: After capture with 1 g
conducted. In a first study, only the catalyst was recovered from PEHA in 3 mL water, H2 (80 bar), C-1 (50 μmol), 2-MTHF (10 mL),
the organic layer and reused for successive hydrogenation 145 °C, 72 h. Methanol yields calculated from 1H NMR with Ph−CH3
cycles (see SI). After four cycles, 95% of C-1’s catalytic and Im (A)/ t-BuOH (B) as internal standards for organic and
efficiency of the initial cycle was retained, with a total of 40.5 aqueous layer, respectively. Error in yield calculations ±5%.
mmol of CH3OH formed, demonstrating the high recyclability
of the catalyst, enabled by this biphasic system (Figure 3A). In a reaction and the loss of amine while transferring PEHA
second study (Figure 3B), both the catalyst and capturing solution between glassware.
amine were recovered and reused. 89% of the CO2 capture In conclusion, a tandem system for CO2 capture (even from
efficiency of the amine was retained in the third cycle, along air) in aqueous amine solution and subsequent hydrogenation
with 87% of methanol productivity. The slight loss in capture is to methanol is described where the catalyst and amine can be
most probably due to the presence of formate species after the recycled multiple times without significant loss in effectiveness.
C DOI: 10.1021/jacs.7b12183
J. Am. Chem. Soc. XXXX, XXX, XXX−XXX
Journal of the American Chemical Society Communication

Among the catalysts tested, a well-defined and commercially Klankermayer, J.; Leitner, W. Chem. Sci. 2015, 6, 693−704.
available complex, Ru-MACHO-BH (C-1) was found most (f) Schneidewind, J.; Adam, R.; Baumann, W.; Jackstell, R.; Beller,
effective, whereas, among various amines, high boiling poly- M. Angew. Chem., Int. Ed. 2017, 56, 1890−1893. (g) Li, Y.-N.; Ma, R.;
amine, PEHA, provided the best CH3OH yields. Our next focus He, L.-N.; Diao, Z.-F. Catal. Sci. Technol. 2014, 4, 1498−1512.
(h) Alberico, E.; Nielsen, M. Chem. Commun. 2015, 51, 6714−6725.
in the context of integrated CO2 capture and hydrogenation (i) Sordakis, K.; Tsurusaki, A.; Iguchi, M.; Kawanami, H.; Himeda, Y.;
will be toward developing a continuous CO2 to CH3OH flow Laurenczy, G. Chem. - Eur. J. 2016, 22, 15605−15608.
system.


(7) Rezayee, N. M.; Huff, C. A.; Sanford, M. S. J. Am. Chem. Soc.
2015, 137, 1028−1031.
ASSOCIATED CONTENT (8) Kothandaraman, J.; Goeppert, A.; Czaun, M.; Olah, G. A.;
*
S Supporting Information Prakash, G. K. S. J. Am. Chem. Soc. 2016, 138, 778−781.
The Supporting Information is available free of charge on the (9) (a) Zhang, L.; Han, Z.; Zhao, X.; Wang, Z.; Ding, K. Angew.
ACS Publications website at DOI: 10.1021/jacs.7b12183. Chem., Int. Ed. 2015, 54, 6186−6189. (b) Khusnutdinova, J. R.; Garg,
J. A.; Milstein, D. ACS Catal. 2015, 5, 2416−2422. (c) Kar, S.;
General information and experimental details (PDF) Goeppert, A.; Kothandaraman, J.; Prakash, G. K. S. ACS Catal. 2017,

■ AUTHOR INFORMATION
Corresponding Author
7, 6347−6351. (d) Ribeiro, A. P. C.; Martins, L. M. D. R. S.;
Pombeiro, A. J. L. Green Chem. 2017, 19, 4811−4815. (e) Everett, M.;
Wass, D. F. Chem. Commun. 2017, 53, 9502−9504. (f) Sordakis, K.;
Tang, C.; Vogt, L. K.; Junge, H.; Dyson, P. J.; Beller, M.; Laurenczy, G.
*gprakash@usc.edu Chem. Rev. 2017, DOI: 10.1021/acs.chemrev.7b00182. (g) Kar, S.;
ORCID Kothandaraman, J.; Goeppert, A.; Prakash, G. K. S. J. CO2 Util. 2018,
G. K. Surya Prakash: 0000-0002-6350-8325 23, 212−218.
(10) (a) Rochelle, G. T. Science 2009, 325, 1652−1654.
Notes
(b) Oyenekan, B. A.; Rochelle, G. T. Ind. Eng. Chem. Res. 2006, 45,
The authors declare no competing financial interest.


2457−2464. (c) Bonenfant, D.; Mimeault, M.; Hausler, R. Ind. Eng.
Chem. Res. 2003, 42, 3179−3184. (d) Yang, H.; Xu, Z.; Fan, M.;
ACKNOWLEDGMENTS Gupta, R.; Slimane, R. B.; Bland, A. E.; Wright, I. J. Environ. Sci. 2008,
Support of our work by the Loker Hydrocarbon Research 20, 14−27. (e) Aaron, D.; Tsouris, C. Sep. Sci. Technol. 2005, 40, 321−
Institute, USC is gratefully acknowledged. 348. (f) Yu, C.-H.; Huang, C.-H.; Tan, C.-S. Aerosol Air Qual. Res.


2012, 12, 745−769.
REFERENCES (11) (a) Kothandaraman, J.; Goeppert, A.; Czaun, M.; Olah, G. A.;
Surya Prakash, G. K. Green Chem. 2016, 18, 5831−5838. (b) Scott, M.;
(1) (a) Macdowell, N.; Florin, N.; Buchard, A.; Hallett, J.; Galindo, Blas Molinos, B.; Westhues, C.; Franciò, G.; Leitner, W. ChemSusChem
A.; Jackson, G.; Adjiman, C. S.; Williams, C. K.; Shah, N.; Fennell, P. 2017, 10, 1085−1093.
Energy Environ. Sci. 2010, 3, 1645−1669. (b) House, K. Z.; Baclig, A.
C.; Ranjan, M.; van Nierop, E. A.; Wilcox, J.; Herzog, H. J. Proc. Natl.
Acad. Sci. U. S. A. 2011, 108, 20428−20433. (c) Goeppert, A.; Czaun,
M.; Prakash, G. K. S.; Olah, G. A. Energy Environ. Sci. 2012, 5, 7833−
7853. (d) Lackner, K. S.; Brennan, S.; Matter, J. M.; Park, A.-H. A.;
Wright, A.; van der Zwaan, B. Proc. Natl. Acad. Sci. U. S. A. 2012, 109,
13156−13162. (e) Sanz-Pérez, E. S.; Murdock, C. R.; Didas, S. A.;
Jones, C. W. Chem. Rev. 2016, 116, 11840−11876.
(2) (a) Aresta, M.; Dibenedetto, A. Dalton Trans. 2007, 2975−2992.
(b) Dibenedetto, A.; Angelini, A.; Stufano, P. J. Chem. Technol.
Biotechnol. 2014, 89, 334−353.
(3) (a) Olah, G. A.; Goeppert, A.; Prakash, G. K. S. Beyond Oil and
Gas: The Methanol Economy, 2nd ed.; Wiley-VCH: Weinheim,
Germany, 2009. (b) Olah, G. A. Angew. Chem., Int. Ed. 2005, 44,
2636. (c) Olah, G. A.; Goeppert, A.; Prakash, G. K. S. J. Org. Chem.
2009, 74, 487. (d) Olah, G. A.; Prakash, G. K. S.; Goeppert, A. J. Am.
Chem. Soc. 2011, 133, 12881−12898. (e) Goeppert, A.; Czaun, M.;
Jones, J.-P.; Prakash, G. K. S.; Olah, G. A. Chem. Soc. Rev. 2014, 43,
7995−8048. (f) Natte, K.; Neumann, H.; Beller, M.; Jagadeesh, R. V.
Angew. Chem., Int. Ed. 2017, 56, 6384−6394.
(4) Obama, B. Science 2017, 355, 126−129.
(5) (a) Zhang, Y.; Fei, J.; Yu, Y.; Zheng, X. Energy Convers. Manage.
2006, 47, 3360−3367. (b) Graciani, J.; Mudiyanselage, K.; Xu, F.;
Baber, A. E.; Evans, J.; Senanayake, S. D.; Stacchiola, D. J.; Liu, P.;
Hrbek, J.; Sanz, J. F.; Rodriguez, J. A. Science 2014, 345, 546−550.
(c) Liu, C.; Yang, B.; Tyo, E.; Seifert, S.; DeBartolo, J.; von Issendorff,
B.; Zapol, P.; Vajda, S.; Curtiss, L. A. J. Am. Chem. Soc. 2015, 137,
8676−8679.
(6) (a) Balaraman, E.; Gunanathan, C.; Zhang, J.; Shimon, L. J. W.;
Milstein, D. Nat. Chem. 2011, 3, 609−614. (b) Han, Z.; Rong, L.; Wu,
J.; Zhang, L.; Wang, Z.; Ding, K. Angew. Chem., Int. Ed. 2012, 51,
13041−13045. (c) Huff, C. A.; Sanford, M. S. J. Am. Chem. Soc. 2011,
133, 18122−18125. (d) Wesselbaum, S.; Vom Stein, T.;
Klankermayer, J.; Leitner, W. Angew. Chem., Int. Ed. 2012, 51,
7499−7502. (e) Wesselbaum, S.; Moha, V.; Meuresch, M.; Brosinski,
S.; Thenert, K. M.; Kothe, J.; Stein, T. v.; Englert, U.; Hölscher, M.;

D DOI: 10.1021/jacs.7b12183
J. Am. Chem. Soc. XXXX, XXX, XXX−XXX

You might also like