You are on page 1of 5

DOI: 10.1002/chem.

201404480 Communication

& Self-Assembly | Hot Paper |

Functionalized Membranes for Photocatalytic Hydrogen


Production
Stefan Troppmann and Burkhard Kçnig*[a]

Chem. Eur. J. 2014, 20, 14570 – 14574 14570  2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Communication

Abstract: Functionalized vesicles for photocatalytic hydro-


gen production in water have been prepared by co-em-
bedding of amphiphilic photosensitizers and a hydrogen-
evolving catalyst in phospholipid membranes. The self-as-
sembly allows a simple two-dimensional arrangement of
the multicomponent system with close spatial proximity,
which gave turnover numbers up to 165 for the incorpo-
rated amphiphilic cobaloxime water reduction catalyst 3 b
under optimized conditions in purely aqueous solution.
Superior photocatalytic activity in fluid membranes indi-
cates that mobility and dynamic reorganization of catalytic
subunits in the membrane promote the visible-light-
driven hydrogen production. The functionalized mem-
branes represent nanostructured assemblies for hydrogen
Figure 1. Self-assembled functionalized membranes for photocatalytic hy-
production in aqueous solution mimicking natural photo-
drogen generation by co-embedding of ruthenium photosensitizer 1 and
synthesis. hydrogen evolving catalyst 3 b in phospholipid membranes.

The conversion of sunlight energy into chemical energy by ous solution without the need of any organic solvent. The in-
means of photocatalytic production of hydrogen is one of the corporation of the two redox-active subunits into lipid bilayers
most promising approaches towards a sufficient and sustaina- leads to a two-dimensional arrangement resulting in a high
ble energy supply, but still a challenging task.[1] Photochemical local concentration of the embedded photosensitizer and cata-
hydrogen generation typically uses three-component systems lyst at the surface. The structured assembly put the complexes
consisting of a light-absorbing photosensitizer, a water-reduc- in close spatial proximity without direct covalent linkage, and
tion catalyst, and a sacrificial electron donor. The electron membrane fluidity allows the dynamic reorganization, both of
transfer between the photosensitizer and the reduction cata- which may accelerate efficient electron transfer and improve
lyst can be enhanced by covalent connection or assembly the catalytic activity.
through bridging ligands leading to a more efficient hydrogen For incorporation into vesicular membranes, the amphiphilic
generation. In contrast, the intermolecular electron transfer in photosensitizers 1, 2 b, and the amphiphilic cobalt catalyst 3 b
multicomponent systems with isolated subunits is diffusion were synthesized (Figure 2). The amphiphilic derivative 1 of
limited. However, their synthesis avoids bridging ligands, is
much simpler and allows varying composition and ratio of
photosensitizer and catalyst. A variety of single- and multicom-
ponent systems for photocatalytic hydrogen generation have
been developed,[2] and many of them use cobalt reduction cat-
alysts.[3] Although water is the ideal medium for H2 production,
almost all reported systems with Co catalysts use organic sol-
vents or mixtures of organic solvents and water. It was demon-
strated that increasing the water content resulted in a de-
creased catalytic activity,[4] and cobalt-based photocatalytic sys-
tems working in purely aqueous solution are rare.[5]
Herein, we report the photocatalytic hydrogen generation in
pure water using self-assembled functionalized vesicle mem-
branes by co-embedding of amphiphilic photosensitizers in
Figure 2. Structures of photosensitizers (1, 2 a, and 2 b) and hydrogen-evolv-
combination with an amphiphilic cobalt-based hydrogen- ing catalysts (3 a and 3 b) used for photocatalytic hydrogen production.
evolving catalyst (Figure 1). Vesicle membranes have been re-
ported as model systems to investigate electron transfer across
membranes and energy conversion.[6] Using vesicles overcomes the established [Ru(bpy)3]2 + (bpy = 2,2’-bipyridine) photosensi-
solubility problems and allows hydrogen production in aque- tizer was prepared by modification of one bipyridine ligand
with long alkyl chains.[7] As metal-free alternative, the xanthene
[a] S. Troppmann, Prof. Dr. B. Kçnig photosensitizer 2 b was tested for light absorption. Catalyst 3 b
Institute of Organic Chemistry for the photochemical proton reduction in vesicular mem-
University of Regensburg, 93040 Regensburg (Germany)
branes was obtained by attaching a hydrophobic hydrocarbon
Fax: (+ 49) 941-943-1717
E-mail: burkhard.koenig@ur.de chain on the axial pyridine ligand of the literature-reported
Supporting information for this article is available on the WWW under CoIII complex [Co(dmgH)2(py)Cl] 3 a (dmgH = dimethylglyoxi-
http://dx.doi.org/10.1002/chem.201404480. mate).[8]

Chem. Eur. J. 2014, 20, 14570 – 14574 www.chemeurj.org 14571  2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Communication

Functionalized vesicles for photocatalytic hydrogen produc- 15 %, respectively (Figure S5 in the Supporting Information). In
tion were prepared by sonication of a lipid film comprising the absence of TEOA, no hydrogen could be quantified by
a phospholipid (4–8, Table 2), either photosensitizer 1 or 2 b head-space gas chromatography.[10]
and catalyst 3 b in aqueous solution. Clear vesicular solutions As was already reported, the hydrogen evolution by using
were obtained, and dynamic light scattering (DLS) analysis cobaloxime complexes as reduction catalysts depends on the
showed small unilamellar vesicles (Figure S1 in the Supporting stability of these complexes, which can be enhanced by the
Information). The incorporation of the amphiphilic photosensi- addition of free dimethylglyoxime ligand to the solution.[9a]
tizer and catalyst into vesicular membranes was proven by Without free dmgH2, TONs of 23 and 93 for 3 b and 1, respec-
size-exclusion chromatography (Figures S2 and S3 in the Sup- tively, were achieved. In the presence of 32 equivalents of free
porting Information), and the stability of vesicles was checked dmgH2 (vs. catalyst 3 b), the catalytic activity increased to 40
by DLS measurements (Figure S4 in the Supporting Informa- and 159, respectively (Figure S6 in the Supporting Informa-
tion). The samples were degassed with nitrogen and irradiated tion).
at constant temperature with high-power light-emitting diodes Next, the influence of the ratio of the two redox-active subu-
(LEDs). The amount of evolved hydrogen was quantified by nits was investigated. Performing the photochemical hydrogen
head-space gas chromatography. For detailed experimental generation in a 1:1 ratio of catalyst 3 b and photosensitizer
setup and data, see the Supporting Information. 1 (126 mm of each component) embedded in DOPC (4) result-
Figure 3 shows the effect of the pH on the catalytic hydro- ed in 76 TON (Table 1, entry 3). Decreasing the concentration
gen production upon irradiation of self-assembled 1,2-dioleoyl-

Table 1. Evolved hydrogen with different ratios of catalyst (cat.) 3 b and


photosensitizer (PS) 1 co-embedded in DOPC (4) in aqueous solution,
15 % TEOA, 4 mm dmgH2, pH 8.3 and irradiation over 13 h at 19 8C.

Entry Ratio cat. ccat3b cPS1 cDOPC n (H2) TON cat. TON
3 b/PS1 [mm] [mm] [mm] [mmol] 3b PS1
1 8:1 126 15.8 858 4.6 9 72
2 4:1 126 31.5 843 20.0 40 159
3 1:1 126 126 748 38.4 76 76
4 1:8 15.8 126 858 7.3 115 14
5 1:32 3.9 126 866 2.6 165 5

of photosensitizer 1 to 31.5 mm gave an increased hydrogen


evolution of 159 turnovers for 1 (entry 2), whereas a loss of ac-
tivity was observed at lower concentrations of 1 (entry 1). Simi-
larly, a decrease in the concentration of catalyst 3 b to 15.8 or
3.9 mm at a constant concentration of 126 mm for photosensi-
tizer 1 resulted in higher TONs of 115 and 165 for the catalyst
3 b, respectively (entries 4 and 5).[11] These results indicate that
Figure 3. Influence of the pH on photocatalytic hydrogen generation from
a system composed of 1 (31.5 mm) and 3 b (126 mm) embedded in DOPC (4,
the high local concentration of co-embedded catalytic sub-
843 mm) in 15 % TEOA aqueous solution after 13 h of irradiation at 19 8C. units allows generation of hydrogen at comparable low cata-
lyst concentration with high catalytic activity.
Control experiments confirmed that every single component
sn-glycero-3-phosphocholine (DOPC, 4) vesicles containing of the photocatalytic system is necessary. No hydrogen pro-
12.6 mol % of 3 b and 3.15 mol % of 1 in a 15 % triethanol- duction could be observed in the absence of either photosen-
amine (TEOA) aqueous solution. The highest activity was ob- sitizer 1, reducing catalyst 3 b, or light.
served at pH 8.3 with a maximum amount of H2 of 11.8 mmol, To investigate the influence of the synthetic phospholipid
resulting in a turnover number (TON) of 23 and 93 based on on the catalytic activity, we co-embedded photosensitizer
3 b and 1, respectively. The catalytic activity dropped with in- 1 (126 mm) and catalyst 3 b (31.5 mm) in vesicles prepared from
creasing or decreasing the pH value. At more basic conditions, different lipids, DOPC (4), 1,2-dimyristoyl-sn-glycero-3-phospho-
the protonation of the reduced CoI catalyst is restricted, thus choline (DMPC, 5), 1-stearoyl-2-myristoyl-sn-glycero-3-phospho-
limiting hydrogen evolution.[9] At more acidic pH, the low con- choline (SMPC, 6), 1,2-dipalmitoyl-sn-glycero-3-phosphocholine
centration of nonprotonated electron donor TEOA reduces the (DPPC, 7), and 1,2-distearoyl-sn-glycero-3-phosphocholine
catalytic activity. (DSPC, 8). The different phospholipids have the same hydro-
Investigating different electron donors, TEOA was found to philic head group but different hydrophobic tails, changing
be most efficient. Upon using TEOA as sacrificial electron the membrane properties significantly (Table 2). We observed
donor, the turnover number for 1 rose from 50 to 93 with in- a strong influence of the nature of the phospholipid and the
creasing the concentration of the tertiary amine from 5 to membrane fluidity on the hydrogen production. The highest

Chem. Eur. J. 2014, 20, 14570 – 14574 www.chemeurj.org 14572  2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Communication

DOPC (4). Nevertheless, the overall performance also depends


Table 2. Chemical structure and phase-transition temperature of different
phospholipids (4–8) used for membrane preparation (Tm is phase-transi- on the nature of the lipid. For SMPC (6) and DSPC (8) vesicles
tion temperature). an increase in the TON above the transition temperature was
observed, but the overall performance was comparatively low.
To get an insight into the mechanism of the photoinduced
intermolecular electron transfer, the redox potentials of photo-
sensitizer 1 and catalyst 3 b were measured (Figure S8 in the
R1 R2 Tm [8C] Supporting Information). The redox potentials indicate that ox-
idative quenching of the excited photosensitizer 1 is feasible
DOPC (4) D9-cis-C17H35 D9-cis-C17H35 21
DMPC (5) C13H27 C13H27 24 from a thermodynamic point of view. The calculated free
SMPC (6) C17H35 C13H27 30 energy DG for the oxidation of the excited photosensitizer
DPPC (7) C15H31 C15H31 41 1 by catalyst 3 b is slightly negative, and the emission quench-
DSPC (8) C17H35 C17H35 55
ing of 1 upon addition of 3 b confirms the possibility of an oxi-
dative quenching (Figures S9 and S10 in the Supporting Infor-
mation). However, the reductive quenching of the excited pho-
tosensitizer 1 by the sacrificial electron donor TEOA is expect-
ed to be the major pathway. The concentration of the electron
donor in solution is much higher than the catalyst concentra-
tion, and reductive quenching yields RuI, which reduces CoII to
CoI inducing hydrogen generation upon formation of a CoIII hy-
dride.
The concept of co-embedding was extended to the amphi-
philic EosinY photosensitizer 2 b in an attempt to reduce the
metal content of the catalytic system. Vesicular systems con-
sisting of DOPC (4), photosensitizer 2 b (126 mm), catalyst 3 b
(126 mm), and 4 mm dmgH2 in 10 % TEOA aqueous solution
pH 7.5 gave 13 TON after 18 h of light irradiation with 535 nm.
Thus, the TON under optimized conditions by using 2 b as
photosensitizer decreased to one-sixth compared with photo-
Figure 4. Effect of the phospholipid (4–8) and the reaction temperature on sensitizer 1 (76 TON) for a 1:1 ratio of the catalytic subunits. A
the photocatalytic hydrogen generation with co-embedded chromophore significant decrease in the fluorescence intensity for embedded
1 (31.5 mm) and catalyst 3 b (126 mm), 4 mm dmgH2 and 15 % TEOA aqueous
photosensitizer 2 b compared to homogeneous solution indi-
solution at pH 8.3 upon irradiation for 12 h.
cates clustering and self-quenching of the xanthene dye 2 b in
the membrane (Figure S11 in the Supporting Information). We
hydrogen generation activity at 19 8C was observed with a turn- assume that the close proximity of 2 b in the lipid bilayer leads
over number of 159 for chromophore 1 by using DOPC (4; to stacking of the photosensitizer in patches preventing an ef-
Figure 4, for kinetic measurements, see Figure S7 in the Sup- ficient intermolecular electron transfer to the catalyst. TONs up
porting Information). Replacing the unsaturated DOPC (4) by to 251 for a homogeneous system by using 2 b and 3 a were
saturated phospholipids (5–8), the catalytic performance of the comparable to literature-reported hydrogen production based
vesicular systems decreased significantly. In DMPC (5), SMPC on EosinY 2 a and 3 a,[9a] thus excluding the synthetic modifica-
(6), DPPC (7), and DSPC (8), the TON dropped to 91, 42, 31, tion of 2 a as a reason for low hydrogen production (Table S1
and 20, respectively. We explain this by the fluidity of the in the Supporting Information).
membrane and the mobility of the embedded photosensitizer In conclusion, self-assembled functionalized vesicles for pho-
1 and water-reduction catalyst 3 b in the membrane. In DOPC tocatalytic hydrogen production in purely aqueous solution
(4) vesicles, with a gel to liquid-crystalline phase transition have been developed. The membrane co-embedded amphi-
temperature (Tm) of 21 8C, the membrane in its liquid-crystal- philic ruthenium photosensitizer and cobalt-based water-re-
line state allows the diffusion of the embedded catalytic sub- duction catalyst evolve hydrogen upon irradiation in the pres-
units in the fluid membrane. The dynamic self-organization re- ence of TEOA as sacrificial electron donor giving TONs up to
sults in high activity with a beneficial distribution of photosen- 165 for the cobalt catalyst under optimal conditions in water.
sitizer and catalyst in the membrane. Vesicles made from satu- Membrane fluidity affects the assembly structure of the incor-
rated phospholipids (5–8) have a higher phase-transition tem- porated catalytic subunits and allows the control of the photo-
perature and the diffusion in the rigid gel membrane is induced hydrogen generation. Best catalytic activity was ob-
restricted at 19 8C, thus, the hydrogen production decreases. served in fluid membranes with high mobility of the photosen-
However, the H2 evolution enhanced for each phospholipid sitizer and the catalyst. Embedding of an amphiphilic EosinY as
above the phase-transition temperature of the respective photosensitizer reduced the metal content of the system but
phospholipid. For DMPC (5) and DPPC (7) in the liquid-crystal- resulted in a significant self-quenching of the chromophores
line state, the overall hydrogen evolution was comparable to with decreased hydrogen production. The membrane embed-

Chem. Eur. J. 2014, 20, 14570 – 14574 www.chemeurj.org 14573  2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Communication

ding allows the functionalization of soft surfaces and may pro- 124; i) R. S. Khnayzer, V. S. Thoi, M. Nippe, A. E. King, J. W. Jurss, K. A.
vide an approach for overall water splitting with membrane El Roz, J. R. Long, C. J. Chang, F. N. Castellano, Energy Environ. Sci. 2014,
7, 1477 – 1488; j) W. M. Singh, T. Baine, S. Kudo, S. Tian, X. A. N. Ma, H.
separated half reactions and transmembrane electron transfer Zhou, N. J. DeYonker, T. C. Pham, J. C. Bollinger, D. L. Baker, B. Yan, C. E.
mimicking natural photosynthesis. Webster, X. Zhao, Angew. Chem. Int. Ed. 2012, 51, 5941 – 5944; Angew.
Chem. 2012, 124, 6043 – 6046; k) B. Shan, T. Baine, X. A. N. Ma, X. Zhao,
R. H. Schmehl, Inorg. Chem. 2013, 52, 4853 – 4859; l) S. Varma, C. E. Cas-
Acknowledgements tillo, T. Stoll, J. Fortage, A. G. Blackman, F. Molton, A. Deronzier, M.-N.
Collomb, Phys. Chem. Chem. Phys. 2013, 15, 17544 – 17552; m) C. Gim-
bert-SuriÇach, J. Albero, T. Stoll, J. Fortage, M.-N. Collomb, A. Deronzier,
We thank the German Science Foundation (DFG) for financial
E. Palomares, A. Llobet, J. Am. Chem. Soc. 2014, 136, 7655 – 7661; n) F.
support. Lakadamyali, E. Reisner, Chem. Commun. 2011, 47, 1695 – 1697; o) L.
Tong, R. Zong, R. P. Thummel, J. Am. Chem. Soc. 2014, 136, 4881 – 4884;
p) M. Natali, A. Luisa, E. Iengo, F. Scandola, Chem. Commun. 2014, 50,
Keywords: photocatalysis · photosynthesis · self-assembly · 1842 – 1844; q) J. Xie, Q. Zhou, C. Li, W. Wang, Y. Hou, B. Zhang, X.
supramolecular chemistry · vesicles Wang, Chem. Commun. 2014, 50, 6520 – 6522.
[6] a) T. Mizushima, A. Yoshida, A. Harada, Y. Yoneda, T. Minatani, S. Murata,
[1] a) N. S. Lewis, D. G. Nocera, Proc. Natl. Acad. Sci. USA 2006, 103, 15729 – Org. Biomol. Chem. 2006, 4, 4336 – 4344; b) S. Bhosale, A. L. Sisson, P. Ta-
15735; b) T. R. Cook, D. K. Dogutan, S. Y. Reece, Y. Surendranath, T. S. lukdar, A. Frstenberg, N. Banerji, E. Vauthey, G. Bollot, J. Mareda, C.
Teets, D. G. Nocera, Chem. Rev. 2010, 110, 6474 – 6502; c) R. Eisenberg, Rçger, F. Wrthner, N. Sakai, S. Matile, Science 2006, 313, 84 – 86; c) A.
D. G. Nocera, Inorg. Chem. 2005, 44, 6799 – 6801. Perez-Velasco, V. Gorteau, S. Matile, Angew. Chem. Int. Ed. 2008, 47,
[2] a) W. T. Eckenhoff, R. Eisenberg, Dalton Trans. 2012, 41, 13004 – 13021; 921 – 923; Angew. Chem. 2008, 120, 935 – 937; d) G. Steinberg-Yfrach,
b) M. Wang, Y. Na, M. Gorlov, L. Sun, Dalton Trans. 2009, 6458 – 6467; P. A. Liddell, S.-C. Hung, A. L. Moore, D. Gust, T. A. Moore, Nature 1997,
c) E. S. Andreiadis, M. Chavarot-Kerlidou, M. Fontecave, V. Artero, Photo- 385, 239 – 241; e) J. J. Grimaldi, S. Boileau, J.-M. Lehn, Nature 1977, 265,
chem. Photobiol. 2011, 87, 946 – 964; d) L. L. Tinker, N. D. McDaniel, S. 229 – 230; f) K. Watanabe, S.-y. Takizawa, S. Murata, Chem. Lett. 2011, 40,
Bernhard, J. Mater. Chem. 2009, 19, 3328 – 3337. 345 – 347; g) R. F. Khairutdinov, J. K. Hurst, Nature 1999, 402, 509 – 511;
[3] a) V. Artero, M. Chavarot-Kerlidou, M. Fontecave, Angew. Chem. Int. Ed. h) L. Zhu, R. F. Khairutdinov, J. L. Cape, J. K. Hurst, J. Am. Chem. Soc.
2011, 50, 7238 – 7266; Angew. Chem. 2011, 123, 7376 – 7405; b) S. Losse, 2005, 127, 825 – 835; i) D. Hvasanov, J. R. Peterson, P. Thordarson, Chem.
J. G. Vos, S. Rau, Coord. Chem. Rev. 2010, 254, 2492 – 2504; c) J. L. Demp- Sci. 2013, 4, 3833 – 3838; j) J. N. Robinson, D. J. Cole-Hamilton, Chem.
sey, B. S. Brunschwig, J. R. Winkler, H. B. Gray, Acc. Chem. Res. 2009, 42, Soc. Rev. 1991, 20, 49 – 94.
1995 – 2004. [7] M. Hansen, F. Li, L. Sun, B. Konig, Chem. Sci. 2014, 5, 2683 – 2687.
[4] a) J. Hawecker, J. M. Lehn, R. Ziessel, New J. Chem. 1983, 7, 271 – 277; [8] W. C. Trogler, R. C. Stewart, L. A. Epps, L. G. Marzilli, Inorg. Chem. 1974,
b) C. V. Krishnan, B. S. Brunschwig, C. Creutz, N. Sutin, J. Am. Chem. Soc. 13, 1564 – 1570.
1985, 107, 2005 – 2015; c) P. Du, J. Schneider, G. Luo, W. W. Brennessel, [9] a) T. Lazarides, T. McCormick, P. Du, G. Luo, B. Lindley, R. Eisenberg, J.
R. Eisenberg, Inorg. Chem. 2009, 48, 4952 – 4962; d) P. Zhang, M. Wang, Am. Chem. Soc. 2009, 131, 9192 – 9194; b) X. Hu, B. S. Brunschwig, J. C.
J. Dong, X. Li, F. Wang, L. Wu, L. Sun, J. Phys. Chem. C 2010, 114, 15868 – Peters, J. Am. Chem. Soc. 2007, 129, 8988 – 8998.
15874; e) T. M. McCormick, Z. Han, D. J. Weinberg, W. W. Brennessel, P. L. [10] With ascorbic acid (200 mm, pH 2–4) as sacrificial electron donor, just
Holland, R. Eisenberg, Inorg. Chem. 2011, 50, 10660 – 10666; f) A. Call, Z. negligible amounts of hydrogen were produced (TON = 3 based on the
Codol, F. AcuÇa-Pars, J. Lloret-Fillol, Chem. Eur. J. 2014, 20, 6171 – photosensitizer).
6183. [11] A direct comparison of the catalytic activity of our vesicular system
[5] a) G. M. Brown, B. S. Brunschwig, C. Creutz, J. F. Endicott, N. Sutin, J. Am. with literature-reported systems is not possible, because photocatalytic
Chem. Soc. 1979, 101, 1298 – 1300; b) C. V. Krishnan, N. Sutin, J. Am. systems with cobalt water-reduction catalysts working in water use dif-
Chem. Soc. 1981, 103, 2141 – 2142; c) B. Probst, M. Guttentag, A. Roden- ferent cobalt complexes and compositions (see Ref. [5]). Furthermore,
berg, P. Hamm, R. Alberto, Inorg. Chem. 2011, 50, 3404 – 3412; d) M. Gut- the insolubility of the modified photosensitizer and the catalyst in
tentag, A. Rodenberg, R. Kopelent, B. Probst, C. Buchwalder, M. Brand- water did not allow performing the hydrogen generation in water in
sttter, P. Hamm, R. Alberto, Eur. J. Inorg. Chem. 2012, 2012, 59 – 64; the absence of phospholipid.
e) M. Guttentag, A. Rodenberg, C. Bachmann, A. Senn, P. Hamm, R. Al- [12] R. Koynova, M. Caffrey, Biochim. Biophys. Acta Biomembr. 1998, 1376, 91.
berto, Dalton Trans. 2013, 42, 334 – 337; f) C. Bachmann, M. Guttentag,
B. Spingler, R. Alberto, Inorg. Chem. 2013, 52, 6055 – 6061; g) M. Nippe,
R. S. Khnayzer, J. A. Panetier, D. Z. Zee, B. S. Olaiya, M. Head-Gordon,
C. J. Chang, F. N. Castellano, J. R. Long, Chem. Sci. 2013, 4, 3934 – 3945; Received: July 20, 2014
h) Y. Sun, J. Sun, J. R. Long, P. Yang, C. J. Chang, Chem. Sci. 2013, 4, 118 – Published online on October 5, 2014

Chem. Eur. J. 2014, 20, 14570 – 14574 www.chemeurj.org 14574  2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

You might also like