You are on page 1of 6

Angewandte

Communications Chemie

International Edition: DOI: 10.1002/anie.201810945


Hydrogen Storage Materials German Edition: DOI: 10.1002/ange.201810945

Reversible Hydrogen Uptake/Release over a Sodium Phenoxide–


Cyclohexanolate Pair
Yang Yu, Teng He,* Anan Wu,* Qijun Pei, Abhijeet Karkamkar, Tom Autrey, and Ping Chen*

Dedicated to the Dalian Institute of Chemical Physics, Chinese Academy of Sciences on the occasion of its 70th anniversary

Abstract: Hydrogen uptake and release in arene–cycloalkane technological implementation of these hydrocarbons suffers
pairs provide an attractive opportunity for on-board and off- in part from low round-trip storage efficiencies; that is, the
board hydrogen storage. However, the efficiency of arene– large enthalpy change of dehydrogenation (DHd = 60–
cycloalkane pairs currently is limited by unfavorable thermo- 70 kJ molH2@1)[6] requires high temperatures to release hydro-
dynamics for hydrogen release. It is shown here that the gen. Several strategies for circumventing or alleviating the
thermodynamics can be optimized by replacement of H in the thermodynamic constraint have been discussed in the liter-
-OH group of cyclohexanol and phenol with alkali or alkaline ature, such as: 1. Non-thermally driven dehydrogenation via
earth metals. The enthalpy change upon dehydrogenation photocatalysis or electrocatalysis. Li et al. showed encourag-
decreases substantially, which correlates with the delocaliza- ingly that cyclohexane released H2 at ambient temperature
tion of the oxygen electron to the benzene ring in phenoxides. when Pt/TiO2 catalyst was used under visible-light irradia-
Theoretical calculations reveal that replacement of H with tion;[7] dehydrogenation of N-heterocycles can also be
a metal leads to a reduction of the HOMO–LUMO energy gap realized at room temperature through photocatalysis or
and elongation of the C@H bond in the a site in cyclo- electrocatalysis.[8] 2. Compositional alteration to optimize
hexanolate, which indicates that the cyclohexanol is activated the thermodynamic properties of hydrocarbons. Previous
upon metal substitution. The experimental results demonstrate efforts by Pez, Crabtree, and Jessop in tuning the thermody-
that sodium phenoxide–cyclohexanolate, an air- and water- namics of dehydrogenation focused on incorporating heter-
stable pair, can desorb hydrogen at ca. 413 K and 373 K in the oatoms such as N into the aromatic ring[9] or introducing
solid form and in an aqueous solution, respectively. Hydro- electron-donating groups on the aromatic ring.[10] For exam-
genation, on the other hand, is accomplished at temperatures as ple, DHd and Td, defined as the temperature at which the
low as 303 K. Gibbs free energy change for dehydrogenation at 1 bar
hydrogen back pressure equals zero (i.e., DGd = 0) decrease
The development of cost-effective, stable, and efficient from 73.6 kJ molH2@1 and ca. 599 K for the cyclohexane–
energy carriers to store and transport energy, especially that benzene pair to 67.3 kJ molH2@1 and ca. 546 K for the
generated from intermittent and distributed renewable piperidine–pyridine pair, to ca. 65.6 kJ mol H2@1 and ca.
resources, is vitally important. Hydrogen has long been 532 K for the cyclohexylamine–aniline pair.[9b] Density func-
viewed as an ideal energy carrier; however, storing hydrogen tional theory (DFT) calculations also show that increasing the
effectively is considered a grand challenge.[1] Over the past number of N atoms in the ring especially in the 1,3- or 1,3,5-
two decades, significant research effort has focused on arrangement or by incorporating stronger electron-donating
materials development for chemi- and physisorption of capability will reduce DHd effectively.[9b, 10] DHd can be
hydrogen.[2] In particular, storing hydrogen in chemical manipulated by strengthening the aromatic character of the
bonds by the reduction of arenes offers the opportunity for p system to stabilize the arene and/or by weakening C@H or
using mass-produced chemicals[3] such as toluene[4] and N@H bonding to destabilize the cycloalkane (or hetero-
dibenzene toluene[5] as hydrogen carriers to mediate the cycloalkane).
energy storage, transportation, and use chain. However, the

[*] Y. Yu, Dr. T. He, Q. Pei, Prof. P. Chen A. Karkamkar, T. Autrey


Dalian Institute of Chemical Physics, Chinese Academy of Sciences Pacific Northwest National Laboratory
Dalian, 116023 (China) Richland, WA 99352 (USA)
E-mail: heteng@dicp.ac.cn Prof. P. Chen
pchen@dicp.ac.cn State Key Laboratory of Catalysis, Dalian Institute of Chemical
Y. Yu, Q. Pei Physics, Chinese Academy of Sciences
University of Chinese Academy of Sciences Dalian, 116023 (China)
Beijing, 100049 (China) and
Dr. A. Wu Collaborative Innovation Centre of Chemistry for Energy Materials
Fujian Provincial Key Laboratory of Theoretical and Computational (iChEM·2011), Xiamen University
Chemistry, College of Chemistry and Chemical Engineering Fujian, 361005 (China)
Xiamen University Supporting information and the ORCID identification number(s) for
Xiamen, 361005 (China) the author(s) of this article can be found under:
E-mail: ananwu@xmu.edu.cn https://doi.org/10.1002/anie.201810945.

3102 T 2019 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2019, 58, 3102 –3107
Angewandte
Communications Chemie

In this paper, we describe a new approach in materials


development and thermodynamic alteration that involves
introducing alkali and alkaline earth metals into the arene–
cycloalkane pairs. The phenol–cyclohexanol pair (DHd of ca.
64.5 kJ molH2@1 and Td of 516 K[11]) was chosen as the parent
pair that was subsequently converted to alkali (or alkaline
earth) phenoxide–cyclohexanolate pairs. DFT calculations
provide insight into the manipulation of the electronic
properties and thermodynamics of hydrogen storage with
the pairs. Experimental studies show that three equivalents
(equiv.) of hydrogen can be reversibly stored in the sodium
phenoxide–cyclohexanolate pair at 373 K. To the best of our
knowledge, this is the first report on the metalation of the
cycloalkane–arene pair to achieve reversible hydrogen stor-
age under moderate conditions.
The combined annual global production of phenol and
cyclohexanol amounts to 1.1 billion tons. Cycling between
phenol and cyclohexanol with & 6 wt % or & 57 g L@1 rever-
sible hydrogen capacity following Reaction (1) provides an
attractive route for hydrogen storage. Little attention, how-
ever, has been focused on this pair, possibly owing to the
unfavorable DHd, (ca. 64.5 kJ molH2@1 corresponding to Td =
516 K)[11] and side reactions during dehydrogenation etc. On
the other hand, the phenol–cyclohexanol pair has a distinct
feature—the labile H of the OH group allows facile chemical
modification. For example, phenol reacts with NaOH to yield
sodium phenoxide and water as products of the acid–base
reaction.

C6 H5 OH þ 3 H2 Ð C6 H11 OH ð1Þ

We employed the Xls method,[12] which is based on DFT


and neutral networks, to simulate a number of alkali (or
alkaline earth) phenoxides and cyclohexanolates and inves-
tigate the effect of replacing the H of the -OH group with
those basic elements. As shown in Figure 1 a, the calculated
dehydrogenation enthalpy change of cyclohexanol!phenol,
64.5 kJ molH2@1, is in good agreement with the enthalpy
change derived from known thermodynamic data,[11] which
validates the accuracy of the calculation method used in this
study. Forming phenoxides by reacting phenol with metals
(M = Li, Na, K, Mg, and Ca) is thermodynamically favorable
(Table S1). Analyses of the natural bond orbital charges show
that electrons on O (Figure 1 b and Table S2) of phenoxides
are more delocalized into the arene ring, thus forming p–p
conjugation and consequently stabilizing the phenoxides.
Formation of the corresponding cyclohexanolates is also
exothermic (Table S1), while a general observation is the
elongation of C@H bonding especially at the a site (Figure 1 b
Figure 1. a) Theoretical calculations of gas-phase DHd for metalated
and Table S3). cyclohexanolate– phenoxide pairs as a function of charge in the
As the magnitude of the conjugation effect is significantly benzene ring and the length of C@H bond in the a-site of cyclo-
greater than the hyperconjugation effect of the same type, the hexanolate compared to that of cyclohexanol. The lines are guide to
exothermicity of the conversion of phenol to phenoxide is eye. b) Charge in the benzene ring of phenoxide compared to that of
greater than that of cyclohexanol to cyclohexanolate. The phenol and the lengths of C@H bonds in cyclohexanolate compared to
important consequence is the substantial decrease in DHd of that of cyclohexanol. c) The HOMO–LUMO energy gaps (a.u.) for
cyclohexanol and sodium cyclohexanolate and the interfacial plots of
alkali (alkaline earth) cyclohexanolate–phenoxide pairs (Fig-
the orbitals.
ure 1 a). For example, dehydrogenation of sodium or potas-
sium cyclohexanolate to corresponding phenoxide has DHd
values of 50.4 or 49.3 kJ molH2@1, respectively, which are ca.

Angew. Chem. Int. Ed. 2019, 58, 3102 –3107 T 2019 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.angewandte.org 3103
Angewandte
Communications Chemie

21.9 % and 23.6 % less than that of cyclohexanol to phenol.


Alkali or alkaline earth metals donate more electrons than H
on O, and the electrons delocalize through the p–p conjuga-
tion and stabilize. Therefore, we refer to the charge delocal-
ized from the O atom to the C-ring in phenoxides as an
indication of the metalation effect (Table S2). A nonlinear
relation between the DHd and charge delocalized from the O
atom to the C-ring in phenoxide was obtained (Figure 1 a).
The molecular structures of the Li, Mg, and Ca phenoxides
are nearly C2 symmetric, while Na and K cations bent over the
ring may cause charge redistribution within the system. It is
worth mentioning that the lower DHd calculated here is from
a gas-phase simulation. However, DHd in the solid phase may
be even lower because of the p–p interactions between
phenoxide molecules, which stabilizes the phenoxides more
than cyclohexanolates. We also note the findings of Jessop
and co-workers[10] in which the DHd of cycloalkane scales with
the Hammett parameter (s) of substituting group, where the
Hammett parameter is a measure of the electron-donating
ability of the substituent group. Because the Hammett
parameter of -O@ (@0.81) is more negative than that of -OH
(@0.37),[13] the substantial decrease in DHd for the alkali
(alkaline earth) cyclohexanolate–phenoxide pair also is
understandable from this viewpoint. Because different alkali
or alkaline earth metals may lead to different -Od@ values, the
reduction in DHd would vary depending on the metal
involved. Figure 2. a) X-ray diffraction patterns of synthesized sodium phenoxide
and sodium cyclohexanolate. b) 1H NMR spectra of sodium phenoxide
Because the sodium cyclohexanolate–phenoxide pair is
and sodium cyclohexanolate compared with those of phenol and
more appealing in terms of hydrogen density (4.9 wt %), DHd cyclohexanol in DMSO.
(50.4 kJ molH2@1), and Td (385 K) (DGd = 0, assuming entropy
change in dehydrogenation contributed mainly by hydrogen
release) (see Figure 1 a), we analyzed its electronic properties The synthesized sodium cyclohexanolate, on the other hand,
and compared the results with those of the parent cyclo- has poor crystallinity and presents a set of weak diffraction
hexanol–phenol pair (see Figure 1 c). Evidently, the replace- peaks at 7.688, 16.888, 21.388, and 41.588 that cannot be assigned to
ment of H with Na redistributes the electron density. any known material containing Na, C, O, and H. The 1H NMR
Compared to cyclohexanol, the energy level of highest and FTIR spectra did not show the H resonance of the OH
occupied molecular orbital (HOMO) increases while that of group and the -OH vibration for either cyclohexanolate and
the lowest unoccupied molecular orbital (LUMO) decreases, phenoxide (Figure 2 b and Figure S1). On the other hand, the
13
leading to a narrower HOMO–LUMO energy gap, which C NMR spectrum of phenoxide shows an obvious downshift
indicates sodium cyclohexanolate is easily activated. compared with that of phenol, which is consistent with the
This finding agrees well with the elongated aC@H bond increase in electron density in the C-ring determined from
(Figure 1 b and Table S3) where the first C@H bond dissoci- calculations (Figure S2). All these measurements suggest the
ation is likely to occur. Because of the correlation of the formation of sodium phenoxide and sodium cyclohexanolate
kinetic barrier with the reaction heat (Brønsted–Evans– via Reactions (2) and (3).
Polanyi relations), the increase in C@H bond length at the
a site also scales with the dehydrogenation enthalpy change C6 H5 OH ðsÞ þ NaH ðsÞ ! C6 H5 ONa ðsÞ þ H2 ðgÞ ð2Þ
(Figure 1 a and Table S4). With the substantial decrease in C6 H11 OH ðsÞ þ NaH ðsÞ ! C6 H11 ONa ðsÞ þ H2 ðgÞ ð3Þ
DHd and promising kinetic features, we set out to synthesize
the sodium cyclohexanolate–phenoxide pair and evaluate its To confirm the thermodynamic data (DHd) calculated
hydrogen uptake and release properties. from the Xls method,[12] experiments were designed to
Sodium phenoxide and sodium cyclohexanolate can be measure the enthalpy change for hydrogenation of
synthesized by mechanically ball-milling phenol and cyclo- sodium phenoxide to cyclohexanolate using a C80 Calvet
hexanol, respectively, with NaH accompanied by evolution of calorimeter.[14] The enthalpy change for the hydrogenation
one equiv. hydrogen (details are given in the Experimental measured (Reaction (4)) is approximately @156 kJ mol@1
Section and Supporting Information). We analyzed the (@52 kJ molH2@1), which agrees well with the calculated
products using X-ray diffraction (XRD), nuclear magnetic result (50.4 kJ molH2@1 for dehydrogenation). Details are
resonance (NMR), and Fourier transform infrared (FTIR) given in the Supporting Information.
spectroscopy. The XRD pattern of synthesized sodium
phenoxide matches well with that in the database (Figure 2 a). C6 H5 ONa ðsÞ þ 3 H2 ðgÞ ! C6 H11 ONa ðsÞ ð4Þ

3104 www.angewandte.org T 2019 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2019, 58, 3102 –3107
Angewandte
Communications Chemie

Because of the decrease in DHd, hydrogen release from hydrolyzes into NaOH and cyclohexanol under the conditions
the sodium phenoxide–cyclohexanolate pair would occur applied in the experiment. Under these conditions, hydro-
under moderate conditions. As cyclohexanolate and phenox- genation and dehydrogenation actually occur via Reac-
ide are solid, both were ball-milled with commercial catalysts tion (5), which has a DHd of 182 kJ mol@1 (ca. 61 kJ molH2@1
(5 % Ru/Al2O3 and 5 % Pt/C) to assist the activation of
C6 H5 ONa ðaq:Þ þ H2 O ðlÞ þ 3 H2 ðgÞ
hydrogen and the C@H bond during hydrogenation and ð5Þ
dehydrogenation. As shown in Figure 3 a, hydrogenation of Ð C6 H11 OH ðlÞ þ NaOH ðaq:Þ
sodium phenoxide under 30 bar hydrogen occurred slowly at
ambient temperature in the presence of 5 % Ru/Al2O3 obtained with the C80 calorimeter in the Supporting Infor-
catalyst; ca. 6 equiv. H was absorbed when the temperature mation). However, with a higher DHd, the entropy contribu-
was increased to 150 8C where the product was characterized tion of water may compensate for such an enthalpy loss.
as sodium cyclohexanolate (Figure S9), showing full hydro- As shown in Figure 3 c, ca. 100 % hydrogenation of
genation. The hydrogenated sample then was dehydrogen- sodium phenoxide can be achieved in 7.5 min at 100 8C
ated under vacuum. However, dehydrogenation proceeded under 40 bar of hydrogen and in the presence of commercial
very slowly at 150 8C. We replaced the Ru/Al2O3 catalyst with 5 % Ru/Al2O3 catalyst. Under milder conditions of 30 8C and
5 % Pt/C catalyst and obtained appropriate dehydrogenation 2 bar hydrogen, hydrogenation of sodium phenoxide also can
at temperatures above 100 8C; about 5 equiv. H was desorbed be achieved within 90 minutes. Cyclohexanol was detected by
1
at 140 8C (Figure 3 b), where the starting dehydrogenation H NMR measurements (Figure S11). Previous studies also
temperature is around 90 8C and sodium phenoxide was showed that the hydrogenation of hydrocarbons such as
obtained finally (Figure S10). Therefore, the metalation phenol[15] and toluene[16] can be carried out facilely under mild
strategy proposed here successfully decreased DHd through conditions. However, dehydrogenation is usually kinetically
forming the phenoxide–cyclohexanolate pair, leading to problematic. Therefore, much attention is given to dehydro-
lower Td. genation.
The slow kinetics in hydrogen uptake and release at lower Because cyclohexanol and NaOH are the hydrogenation
temperatures shown in Figure 3 a,b implies the inefficient products, we conducted the dehydrogenation of cyclohexanol
catalysis of solid Ru/Al2O3 or Pt/C in solid-state reactions. and NaOH in aqueous solution catalyzed by commercial 5 %
Therefore, we dissolved the reactant and catalyst in water and Pt/C. Both phenoxide and cyclohexanone were found in the
re-investigated the properties of hydrogenation and dehy- dehydrogenation products as shown in Figure 3 d and Fig-
drogenation. It should be noted that sodium cyclohexanolate ure S12. Selectivity to phenoxide is pH dependent, i.e., the
dehydrogenation of cyclohexanol
in the presence of 1 equiv. of
NaOH (pH 13.3) has a conversion
of cyclohexanol and selectivity to
phenoxide of 56.9 % and 81.0 %
after 20 hours of reaction; with
excess NaOH (pH 14.0); both the
conversion and selectivity reach
60.0 % and 91.7 % within the
same period of time. Prolonging
the reaction leads to > 99 % con-
version of cyclohexanol and selec-
tivity to sodium phenoxide in
aqueous phase. As H2 is the only
detectable gaseous product (Fig-
ure S13), the yield of H2 is > 99 %;
thus, the reverse reaction [Reac-
tion (5)] occurs at a temperature as
low as 100 8C. It is noteworthy that
aqueous sodium phenoxide and
cyclohexanol are stable in air for
weeks (Figure S14), which is
a desirable property for hydrogen
Figure 3. a) Hydrogen uptake by sodium phenoxide catalyzed by 5 % Ru/Al2O3 under 30 bar hydrogen. storage materials.
b) Hydrogen release from sodium cyclohexanolate catalyzed by 5 % Pt/C under vacuum. The molar There are reports demonstrat-
ratios of metals to phenoxide (or cyclohexanolate) are 1:10. c) Hydrogenation of sodium phenoxide ing that metalated (by alkali or
in aqueous solution catalyzed by 5 % Ru/Al2O3 with a molar ratio of Ru to phenoxide of 1:30. *:
alkaline earth metals) amidobor-
Cphenoxide = 0.148 mol L@1, molar ratio of Ru to phenoxide is 1:15. d) Dehydrogenation of cyclohexanol
in NaOH aqueous solution catalyzed by 5 % Pt/Al2O3 at different pH values for 20 h or 72 h with anes,[17] hydrazinoboranes,[18] and
a molar ratio of Pt to cyclohexanol of 1:50. Detailed information can be found in the Supporting primary amines[19] have lower
Information. dehydrogenation temperatures

Angew. Chem. Int. Ed. 2019, 58, 3102 –3107 T 2019 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.angewandte.org 3105
Angewandte
Communications Chemie

and enhanced selectivity to hydrogen in comparison to their document-onboard-hydrogen-storage-light-duty-fuel-cell; d) J.


neat forms. In the present work, such a prominent effect of Yang, A. Sudik, C. Wolverton, D. J. Siegel, Chem. Soc. Rev.
metalation also was achieved. The sodium-modified cyclo- 2010, 39, 656 – 675; e) U. Eberle, M. Felderhoff, F. Schgth,
Angew. Chem. Int. Ed. 2009, 48, 6608 – 6630; Angew. Chem.
hexanol–phenol pair can undergo dehydrogenation at 100 8C,
2009, 121, 6732 – 6757.
which is substantially lower than the Td for cycloalkanes
[2] a) B. Bogdanović, M. Schwickardi, J. Alloys Compd. 1997, 253 –
(generally > 300 8C for thermal dehydrogenation)[6] and for 254, 1 – 9; b) P. Chen, Z. Xiong, J. Luo, J. Lin, K. L. Tan, Nature
N-heterocycles and substituted cycloalkanes (generally 2002, 420, 302 – 304; c) S.-i. Orimo, Y. Nakamori, J. R. Eliseo, A.
> 170 8C for thermal dehydrogenation).[10, 20] Although the Zgttel, C. M. Jensen, Chem. Rev. 2007, 107, 4111 – 4132; d) M.
dehydrogenation of 1,2,3,4-tetrahydroquinolines or indolines Chandra, Q. Xu, J. Power Sources 2006, 156, 190 – 194; e) D.
can be achieved at relatively low temperature (refluxing of Teichmann, K. Stark, K. Muller, G. Zottl, P. Wasserscheid, W.
solvents such as toluene, xylene, or 2,2,2-trifluoroethanol),[21] Arlt, Energy Environ. Sci 2012, 5, 9044 – 9054; f) S. Ott, Science
only the N-containing rings were involved in the hydrogen 2011, 333, 1714 – 1715; g) N. L. Rosi, J. Eckert, M. Eddaoudi,
D. T. Vodak, J. Kim, M. OQKeeffe, O. M. Yaghi, Science 2003,
release due to the delocalized p bonding in the whole
300, 1127 – 1129; h) M. P. Suh, H. J. Park, T. K. Prasad, D.-W.
molecule after dehydrogenation. Furthermore, sodium phen- Lim, Chem. Rev. 2012, 112, 782 – 835.
oxide can be prepared by reacting phenol and NaOH; [3] a) Q.-L. Zhu, Q. Xu, Energy Environ. Sci. 2015, 8, 478 – 512; b) P.
therefore, it can be mass produced at low cost. It is stable in Preuster, C. Papp, P. Wasserscheid, Acc. Chem. Res. 2017, 50,
air and water and has a low vapor pressure. All these 74 – 85.
characteristics make the sodium phenoxide–cyclohexanolate [4] A. Shukla, S. Karmakar, R. B. Biniwale, Int. J. Hydrogen Energy
pair a very attractive hydrogen storage material for practical 2012, 37, 3719 – 3726.
applications. Further efforts should focus on materials [5] a) H. Jorschick, P. Preuster, S. Durr, A. Seidel, K. Muller, A.
Bosmann, P. Wasserscheid, Energy Environ. Sci. 2017, 10, 1652 –
engineering including developing more efficient hydrogena-
1659; b) D. Geburtig, P. Preuster, A. Bçsmann, K. Mgller, P.
tion/dehydrogenation catalysts and reaction media. Wasserscheid, Int. J. Hydrogen Energy 2016, 41, 1010 – 1017.
[6] R. B. Biniwale, S. Rayalu, S. Devotta, M. Ichikawa, Int. J.
Hydrogen Energy 2008, 33, 360 – 365.
Acknowledgements [7] L. Li, X. Mu, W. Liu, Z. Mi, C.-J. Li, J. Am. Chem. Soc. 2015, 137,
7576 – 7579.
T.H. and P.C. acknowledge support provided by the National [8] a) K.-H. He, F.-F. Tan, C.-Z. Zhou, G.-J. Zhou, X.-L. Yang, Y. Li,
Natural Science Foundation of China (21875246, 51671178, Angew. Chem. Int. Ed. 2017, 56, 3080 – 3084; Angew. Chem.
2017, 129, 3126 – 3130; b) S. Kato, Y. Saga, M. Kojima, H. Fuse, S.
51472237), DICP (DICP ZZBS201616), Sino-Japanese
Matsunaga, A. Fukatsu, M. Kondo, S. Masaoka, M. Kanai, J.
Research Cooperative Program of the Ministry of Science
Am. Chem. Soc. 2017, 139, 2204 – 2207; c) Y. Wu, H. Yi, A. Lei,
and Technology (2016YFE0118300), and iChEM·2011. A.W. ACS Catal. 2018, 8, 1192 – 1196; d) Q. Yin, M. Oestreich, Angew.
acknowledges the financial support provided by the National Chem. Int. Ed. 2017, 56, 7716 – 7718; Angew. Chem. 2017, 129,
Science Foundation of China (NSFC) (21773193) and the 7824 – 7826; e) M. Zheng, J. Shi, T. Yuan, X. Wang, Angew.
Fundamental Research Funds for the Central Universities Chem. Int. Ed. 2018, 57, 5487 – 5491; Angew. Chem. 2018, 130,
(20720160031). T.A. and A.K. gratefully acknowledge sup- 5585 – 5589.
port from the Hydrogen Materials—Advanced Research [9] a) G. P. Pez, A. R. Scott, A. C. Cooper, H. Cheng, F. C. Wilhelm,
Consortium (HyMARC), established as part of the Energy A. H. Abdourazak, US7351395B1, 2008; b) E. Clot, O. Eisen-
stein, R. H. Crabtree, Chem. Commun. 2007, 2231 – 2233;
Materials Network under the U.S. Department of Energy,
c) R. H. Crabtree, Energy Environ. Sci. 2008, 1, 134 – 138.
Office of Energy Efficiency and Renewable Energy, Fuel Cell [10] Y. Cui, S. Kwok, A. Bucholtz, B. Davis, R. A. Whitney, P. G.
Technologies Office. Pacific Northwest National Laboratory Jessop, New J. Chem. 2008, 32, 1027 – 1037.
is a multi-program national laboratory operated by Battelle [11] Based on gas-phase values of DHf (C6H5OH, g) = @96.4 kJ mol@1
for the U.S. Department of Energy under Contract DE-AC05- and DHf (C6H11OH, g) = @290 kJ mol@1, assuming a similar
76RL01830. entropy change as that of the aniline–cyclohexylamine pair
(125 J mol@1 K).
[12] J. Wu, Z. I. Ying, X. Xu, ChemPhysChem 2010, 11, 2561 –
2567.
Conflict of interest [13] C. Hansch, A. Leo, R. W. Taft, Chem. Rev. 1991, 91, 165 –
195.
The authors declare no conflict of interest. [14] A. Karkamkar, K. Parab, D. M. Camaioni, D. Neiner, H. Cho,
T. K. Nielsen, T. Autrey, Dalton Trans. 2013, 42, 615 – 619.
Keywords: hydrogen storage materials · metalation · [15] D. Forberg, T. Schwob, R. Kempe, Nat. Commun. 2018, 9,
organic hydrides · thermodynamic modification 1751.
[16] M. Zahmakıran, Y. Tonbul, S. :zkar, J. Am. Chem. Soc. 2010,
How to cite: Angew. Chem. Int. Ed. 2019, 58, 3102 – 3107 132, 6541 – 6549.
Angew. Chem. 2019, 131, 3134 – 3139 [17] Z. Xiong, et al., Nat. Mater. 2008, 7, 138 – 141.
[18] H. Wu, W. Zhou, F. E. Pinkerton, T. J. Udovic, T. Yildirim, J. J.
Rush, Energy Environ. Sci. 2012, 5, 7531 – 7535.
[1] a) L. Schlapbach, A. Zuttel, Nature 2001, 414, 353 – 358; b) T. [19] J. Chen, et al., Chem. Eur. J. 2014, 20, 6632 – 6635.
He, P. Pachfule, H. Wu, Q. Xu, P. Chen, Nat. Rev. Mater. 2016, 1, [20] a) Z. Wang, I. Tonks, J. Belli, C. M. Jensen, J. Organomet. Chem.
16059; c) US Department of Energy (DOE) targets, 2009, 694, 2854 – 2857; b) M. Yang, C. Han, G. Ni, J. Wu, H.
https://energy.gov/eere/fuelcells/downloads/target-explanation- Cheng, Int. J. Hydrogen Energy 2012, 37, 12839 – 12845.

3106 www.angewandte.org T 2019 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2019, 58, 3102 –3107
Angewandte
Communications Chemie

[21] a) S. Chakraborty, W. W. Brennessel, W. D. Jones, J. Am. Chem. S. Johnston, M. Yan, J. Xiao, Angew. Chem. Int. Ed. 2013, 52,
Soc. 2014, 136, 8564 – 8567; b) R. Yamaguchi, C. Ikeda, Y. 6983 – 6987; Angew. Chem. 2013, 125, 7121 – 7125.
Takahashi, K.-i. Fujita, J. Am. Chem. Soc. 2009, 131, 8410 – 8412;
c) C. Deraedt, R. Ye, W. T. Ralston, F. D. Toste, G. A. Somorjai, Manuscript received: September 22, 2018
J. Am. Chem. Soc. 2017, 139, 18084 – 18092; d) J. Wu, D. Talwar, Revised manuscript received: November 14, 2018
Accepted manuscript online: November 25, 2018
Version of record online: December 17, 2018

Angew. Chem. Int. Ed. 2019, 58, 3102 –3107 T 2019 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.angewandte.org 3107

You might also like