You are on page 1of 11

Research Article

Chemistry—A European Journal doi.org/10.1002/chem.202302284

www.chemeurj.org

Initial Quenching Efficiency Determines Light-Driven H2


Evolution of [Mo3S13]2 in Lipid Bilayers
Amir Abbas,[a] Eva Oswald,[b] Jan Romer,[b] Anja Lenzer,[a] Magdalena Heiland,[a]
Carsten Streb,[a, c] Christine Kranz,[b] and Andrea Pannwitz*[a]

Nature uses reactive components embedded in biological liposomes provided similar liposomal structures according to
membranes to perform light-driven photosynthesis. Here, a TEM analysis, only DMPC yielded high H2 amounts. In situ
model artificial photosynthetic system for light-driven hydrogen scanning electrochemical microscopy (SECM) measurements
(H2) evolution is reported. The system is based on liposomes using Pd microsensors revealed an induction period of around
where amphiphilic ruthenium trisbipyridine based photosensi- 26 minutes prior to H2 evolution, indicating an activation
tizer (RuC9) and the H2 evolution reaction (HER) catalyst mechanism which might be induced by the fluid-gel phase
[Mo3S13]2 are embedded in biomimetic phospholipid mem- transition of DMPC at room temperature. Stern-Volmer-type
branes. When DMPC was used as the main lipid of these light- quenching studies revealed that electron transfer dynamics
active liposomes, increased catalytic activity (TONCAT ~200) was from the excited state photosensitizer are most efficient in the
observed compared to purely aqueous conditions. Although all DMPC lipid environment giving insight for design of artificial
tested lipid matrixes, including DMPC, DOPG, DPPC and DOPG photosynthetic systems using lipid bilayer membranes.

Introduction harvesting complexes (LHCs), photosystem I and II perform solar


energy absorption and conversion as illustrated in Figure 1. The
The transition to a sustainable energy economy requires the self-assembled lipid bilayer promote (supra)molecular organiza-
development of technology to exploit abundant and renewable tion and high local concentration of reactive components to
resources that can be used to generate green fuels, such as enhance reactivity and suppress side reactions.[5,6] Vesicles with
hydrogen gas (H2).[1] The most promising process – yet still lipid bilayer membranes and embedded functional molecules
challenging – for the H2 production is the light-driven splitting were reported as model systems to perform light absorption
of water by artificial photosynthesis using bioinspired and light-driven reactions such as H2 generation, CO2 reduction
systems.[1–3] Typically, these systems are composed of three and water oxidation by embedding the active units at the
components: a sacrificial electron donor as reductant, a photo- membrane, and thereby overcome organic solvent usage and
sensitizer (PS) and the H2 evolution reaction (HER) catalyst solubility problems.[7–12]
(CAT).[4] In natural photosynthesis, light-driven catalysis is In this work, the light-driven H2 evolution reaction (HER) at
carried out in an aqueous environment, where chromophores lipid bilayer membranes of liposomes is reported, using the
and catalytic units operate within biological membranes. amphiphilic photosensitizer bis(2,2’-bipyridine)-(4,4’-dinonyl-
Specifically, transmembrane protein complexes such as light 2,2’-bipyridine)-ruthenium(II) (RuC9), and the molybdenum
sulfide cluster [Mo3S13]2 as HER catalyst (see Figure 1).
[a] A. Abbas, A. Lenzer, Dr. M. Heiland, Prof. Dr. C. Streb, Jun.- The amphiphilic photosensitizer RuC9 is known to integrate
Prof. Dr. Dr. A. Pannwitz at the bilayer-water interface of phospholipid membranes due
Institute of Inorganic Chemistry I to hydrophobic interactions between the alkyl tails and the
Ulm University hydrophobic core of the membrane, and electrostatic inter-
Albert-Einstein-Allee 11, 89081 Ulm (Germany)
E-mail: andrea.pannwitz@uni-ulm.de actions of its [Ru(bpy)3]2 + -type headgroup and the lipid
[b] E. Oswald, J. Romer, Prof. Dr. C. Kranz headgroups.[5,13–15] Upon incorporation into the membrane,
Institute of Analytical and Bioanalytical Chemistry RuC9 retains its photosensitizer activity, as exemplified for light-
Ulm University driven CO2 reduction.[5,14]
Albert-Einstein-Allee 11, 89081 Ulm (Germany)
Molecular molybdenum sulfides (thiomolybdates), and
[c] Prof. Dr. C. Streb
specifically [Mo3S13]2 , are known to be highly active HER
Department of Chemistry
Johannes Gutenberg University Mainz catalysts when dissolved in organic solvent,[16–19] while their HER
Duesbergweg 10–14, 55128 Mainz (Germany) reactivity is affected in presence of water.[20,21] However, the
Supporting information for this article is available on the WWW under performance of [Mo3S13]2 in heterogenous environments for
https://doi.org/10.1002/chem.202302284 light-driven H2 evolution showed promising activity when using
© 2023 The Authors. Chemistry - A European Journal published by Wiley-VCH soft polymeric materials such as nanoporous polymeric mem-
GmbH. This is an open access article under the terms of the Creative
branes consisting of amphiphilic block copolymers (e. g. P(S-co-
Commons Attribution Non-Commercial NoDerivs License, which permits use
and distribution in any medium, provided the original work is properly cited, I)-b-PDMAEMA),[22] polyampholytic graft copolymers as matrix
the use is non-commercial and no modifications or adaptations are made. for TiO2/Eosin Y/[Mo3S13]2 -organic-inorganic hybrid

Chem. Eur. J. 2023, 29, e202302284 (1 of 11) © 2023 The Authors. Chemistry - A European Journal published by Wiley-VCH GmbH
15213765, 2023, 72, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/chem.202302284 by Indian Institute of Technology Indore, Wiley Online Library on [13/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Research Article
Chemistry—A European Journal doi.org/10.1002/chem.202302284

materials,[23] polymeric carbon nitrides,[24] and polymeric


micelles.[25] Electrostatic attraction between the negatively
charged [Mo3S13]2 and positively charged photosensitizer
molecules such as [Ru(bpy)3]2 + was reported to be beneficial for
immobilization into nanoporous polymeric membranes as well
as charge transfer dynamics in the light-driven HER.[22] Building
on these results, we designed a biomimetic liposome-system
with the positively charged and membrane-bound PS2 + RuC9
that can electrostatically attract the negatively charged HER
catalyst [Mo3S13]2 to perform light-driven H2 evolution. To
investigate the role of the lipid environment, we tested several
lipids with different properties. Besides bulk measurements,
scanning electrochemical microscopy (SECM)[26] was used to
study the onset dynamics upon start of light irradiation and
probe the early phase of the H2 evolution catalysis (methods
see Figure 2).

Results and Discussion

Synthesis and liposomes preparation

The synthesis of the photosensitizer, RuC9,[5,14] and the catalyst,


Na2[Mo3S13] x 5H2O,[21,24,27] was performed as described previ-
ously. Light-active active liposomes were prepared using a
Figure 1. A) Light-driven reactions in natural photosynthesis take place in
the thylakoid membrane of chloroplasts, where electrons are donated via 100 : 1 molar ratio of a selected main lipid with various transition
the electron transfer chain to photosystem I (PSI) which subsequently temperatures and charges (see Table 1) and the sterically
reduces the NADP + reductase for NADH generation. B) H2 evolution in our stabilizing lipid 1,2-dimyristoyl-sn-glycero-3 phosphoethanol-
biomimetic system, using liposomes as membrane for assembling the
charged photosensitizer (PS2 +) RuC9 and the HER catalyst (CAT2 ) [Mo3S13]2 amine-N-[methoxy(polyethylene glycol)-2000] (14:0 PEG2000
along with ascorbic acid as sacrificial electron donor, depicting the electron PE). In a typical experiment, RuC9 and Na2[Mo3S13] x 5H2O were
transfer steps and reaction pathway. Figure A was adapted from “Light incorporated into the lipid bilayer in 100 : 1 : 5 : 1 molar ratio of
Dependent Reactions of Photosynthesis”, by BioRender.com (2022), retrieved
from https://app.biorender.com/biorender-templates. the (main lipid): (14 : 0 PEG2000 PE) : RuC9 : [Mo3S13]2 via a lipid
film rehydration with phosphate buffer (10 mM) and follow-up
extrusion, as described in previous studies and in the Support-
ing Information in Figure S1.[5,28]

Figure 2. Quantification methods used for light-driven HER in this study: A) Bulk liposome samples and sampling of H2 in the gas phase for quantification via
head space gas chromatography. B) SECM setup for in situ electrochemical determination of H2 using a two-compartment electrochemical cell with a track-
etched membrane for separating the liposome containing solution from the electrodes.

Chem. Eur. J. 2023, 29, e202302284 (2 of 11) © 2023 The Authors. Chemistry - A European Journal published by Wiley-VCH GmbH
15213765, 2023, 72, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/chem.202302284 by Indian Institute of Technology Indore, Wiley Online Library on [13/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Research Article
Chemistry—A European Journal doi.org/10.1002/chem.202302284

Table 1. Molecular structures, names and transition temperatures (Tm) of main phospholipids and the sterically stabilizing lipid 14:0 PEG2000 PE.[29]
Phospholipid structure Name Tm Charge of
(°C) the head
group

1,2-dimyristoyl-sn-glycero-3-phosphocholine 14 : 0 PC
1 24 Zwitterionic
(DMPC)

1,2-dioleoyl-sn-glycero-3-phospho-(1’-rac-glycerol)
2 18 Negative
18 : 1 (Δ9-Cis) (DOPG)

1,2-dioleoyl-sn-glycero-3-phosphocholine 18 : 1
3 17 Zwitterionic
(Δ9-Cis) PC (DOPC)

1,2-dipalmitoyl-sn-glycero-3-phosphocholine 16 : 0 PC
4 41 Zwitterionic
(DPPC)

1,2-dimyristoyl-sn-glycero-3 phosphoethanolamine-N-
5 [methoxy(polyethylene glycol)-2000] (14 : 0 PEG2000 n.a. Negative
PE)

Structural properties TEM-analysis confirmed size and structure of the functionalized


vesicles.
The liposomes’ size and morphology were characterized by
dynamic light scattering (DLS) and transmission electron micro-
scopy (TEM). DLS yielded typical hydrodynamic diameters (ZAvg) Light-driven HER
of 140–155 nm (Figure S2), which was confirmed by TEM
imaging (see Figure 3). TEM analysis of light-active liposomes Visible light-driven HER activity was explored by combining the
was carried out via a negative staining technique, using uranyl buffered liposome sample with ascorbic acid as sacrificial
acetate as contrast agent.[30] Figure 3 A–D show TEM images of electron donor at pH 6, (pH was adjusted by diluted aqueous
light-active liposomes based on DMPC, DPPC, DOPC and DOPG. NH4OH solution). The pH was chosen based on efficiency of
These images show mainly wrinkled spherical structures with electron donating properties of ascorbic acid, which is best in
diameters below 200 nm. The wrinkled shapes are typically due its deprotonated form (ascorbate), as previously demonstrated
to the vacuum-drying of the samples that is required for TEM in light-driven H2 evolution reactions of the [Mo3S13]2
analysis.[31] Furthermore, Figure 3 shows some planar structures catalyst.[16,33] Typical light-driven experiments were performed in
that are also generated from the lipid bilayers during the 8 mL GC vials under argon atmosphere at room temperature in
temperature and pressure changes in the TEM experiment a modular 3D printed and fully characterized open-source
leading to structural changes within the vesicles.[32] Overall, the photoreactor,[21] using a 460 nm LED light source and a sample
holder for up to 10 GC vials to be irradiated simultaneously

Figure 3. TEM images of negatively stained liposome samples with a composition of 100 : 1 : 5 : 1 of main lipid:(14 : 0 PEG2000 PE) : RuC9:[Mo3S13]2 , prepared
with 10 mM phosphate buffer and the main lipid being A) DMPC, B) DPPC, C) DOPC and D) DOPG.

Chem. Eur. J. 2023, 29, e202302284 (3 of 11) © 2023 The Authors. Chemistry - A European Journal published by Wiley-VCH GmbH
15213765, 2023, 72, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/chem.202302284 by Indian Institute of Technology Indore, Wiley Online Library on [13/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Research Article
Chemistry—A European Journal doi.org/10.1002/chem.202302284

under identical irradiation conditions (see Figure S3). The which only minor amounts of H2 were produced additionally
photon flux of the setup was 1001 nmol s 1 at the outer glass (see Figure 4A). Such behavior is well known for light-driven
walls of the GC reaction vessels[21] and the reaction mixtures catalysis with [Ru(bpy)3]2 + based photosensitizers and is due to
were stirred during irradiation. Each experiment was carried out decomposition of the photosensitizer during prolonged
in triplicate and quantification of H2 was performed by gas irradiation.[5,34–37] The stability of RuC9 during catalysis was
chromatography of samples from the respective head spaces evaluated via UV-Vis spectroscopy. It was found that the
after 2 h, 4 h, 6 h and 24 h of irradiation (see Figure 4A and characteristic absorption band between 450 and 460 nm almost
Figure S6 for a typical GC data trace). Additionally, a time- disappeared after 24 h of irradiation (Figure S8), indicating that
dependent HER activity of the light-active DMPC vesicles were the RuC9 photosensitizer had been degraded over the course of
performed over 1 h with sampling every 5 minutes (Figure 4B). the light-driven catalysis. H2 production reached a maximum
The performance of the catalyst was quantified via the turnover after around six hours with an average reaction rate of 111
number (TONCAT) which is the amount of H2 produced divided pmol·s 1 which translates to an average turnover frequency of
DTON
by the amount of catalyst according to the following equation: the catalyst of TOFCAT = Dt = 28.5 h 1 over these 6 hours. Under
n ðH Þ
TONCAT = nð ½Mo2 S213 �2 Þ . The TONCAT values in this study always aerated conditions, the HER activity of light-active DMPC
refer to the amount of catalyst which was added during sample liposomes dropped significantly, and after 24 h of irradiation, a
preparation. It is assumed that not all catalyst molecules are TONDMPC,air = 21 � 3 was observed, indicating that oxygen
taken up in the membrane or may be lost in the filter interferes with the productivity of the system. The absence of
membrane during the extrusion of the liposomes. Therefore, phospholipids reduced the TONCAT significantly to TONCAT = 18 �
the real TONCAT may be higher than the herein reported TONCAT. 3, which is in line with previous reports on degradation
The performance of the photosensitizer was also quantified via hydrolysis[16] of the [Mo3S13]2 as well as deactivating
n ðH Þ
its turnover number (TONPS) with TONPS = 2� nð RuC29 Þ . The factor aggregation[38] of positively charged PS with negatively charged
of two arises from the fact that two consecutive one-electron catalyst in water. On the other hand, electrostatic interactions
donation steps are necessary from the PS to the catalyst. As the at soft interfaces, such as previously reported POMbranes can
PS is one-electron donor two turnovers of the PS per molecule promote effective charge transfer between the PS and the
of H2 are required.[10] In previously reported studies, it was catalyst and therefore H2 evolution.[22] As mentioned above,
found that only 75 % of RuC9 was taken up in DPPC vesicles DMPC has zwitterionic head groups, in combination with
under comparable conditions.[14] Based on these data, it can be around 5 % loading of dicationic RuC9, a net-positive surface
estimated that the real TONPS, in all investigated cases, might be charge was expected and confirmed by a zeta potential of
around 30 % higher than we report here, however it should be + 2.3 � 1 mV according to electrophoretic light scattering
noted that we report catalysis not only in DPPC but also in measurements. We therefore anticipate that the positive surface
other lipid environments, where the real uptake might be charge electrostatically attracts the negatively charged
different. [Mo3S13]2 HER catalyst and promotes charge transfer between
It was found that DMPC based liposomes with a composi- the PS and the catalyst, resulting in H2 evolution. The absence
tion of 100 : 1 : 5 : 1 of DMPC:(14 : 0 PEG2000 PE) : RuC9 : [Mo3S13]2 of PS, catalyst, sacrificial electron donor or light resulted in H2
generated H2 with a TONDMPC = 196 � 7 within 24 h. However, production below the detection limit, even after 24 hours of
maximum activity was observed during the first 6 hours after irradiation.

Figure 4. Time-dependent HER activity and TONCAT results of the light-active DMPC vesicles and system where phospholipids were absent, with various
reference experiments A) monitored over 24 h and B) data of the first hour. Experimental conditions: liposome sample with a composition of 100 : 1 : 5 : 1 of
DMPC:(14 : 0 PEG2000 PE) : RuC9 : [Mo3S13]2 , with c(DMPC) = 5 mM, c([Mo3S13]2 ) = 50 μM, pH = 6 and control experiments where phospholipids, ruthenium
photosensitizer, HER catalyst, light source or sacrificial electron donor were absent. All values were acquired in Ar atmosphere and in triplicate with error bars
presenting the standard deviation. Note: The data of the control experiments overlap at 0 μmol and TONCAT = 0 respectively.

Chem. Eur. J. 2023, 29, e202302284 (4 of 11) © 2023 The Authors. Chemistry - A European Journal published by Wiley-VCH GmbH
15213765, 2023, 72, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/chem.202302284 by Indian Institute of Technology Indore, Wiley Online Library on [13/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Research Article
Chemistry—A European Journal doi.org/10.1002/chem.202302284

Table 2. Light-driven H2 evolution (460 nm LED, head space GC sampling) after 24 h of liposome samples with various CAT:PS ratios and ascorbic acid
concentrations. Experimental conditions: liposome samples in Ar atmosphere with a composition of 100 : 1:Y : X of DMPC:(14 : 0 PEG2000 PE) : RuC9 : [Mo3S13]2 ,
with c(DMPC) = 5 mM at various loadings of RuC9 and [Mo3S13]2 and various concentrations of ascorbic acid, respectively. CCAT and CPS indicate the bulk
concentrations.
Entry Cascorbic acid[mM] CAT:PS ratio CCAT [μM] CPS [μM] H2 (μmol) TONCAT TONPS

1 100 1:5 50 250 2.50 � 0.04 196 � 7 78 � 1


2 100 1 : 10 25 250 1.53 244 49
3 100 1 : 20 12.5 250 1.77 566 57
4 100 1 : 40 6.25 250 1.34 856 43
5 100 1:1 50 50 0.43 34.4 69
6 150 1:5 50 250 2.44 � 0.09 195 � 7 78 � 3
7 70 1:5 50 250 2.22 � 0.15 177 � 12 71 � 5
8 40 1:5 50 250 1.07 � 0.06 92 � 4 36 � 2
9 10 1:5 50 250 0.68 � 0.08 54 � 7 22 � 3

Varying the ratio between catalyst and photosensitizer by higher concentrations. Therefore, we investigated higher
reducing the amount of catalyst and keeping RuC9 loading CAT : PS ratios at constant 1 % catalyst loading while reducing
constant reduced the overall H2 evolution (see Table 2) but the photosensitizer loading. As a consequence, we could only
increased the catalyst performance documented by an increas- detect H2 production at 1 : 1 ratio at our setup and not with
ing TONCAT at lower catalyst loading. We recorded the highest lower loadings. At 1 : 1 ratio, we observed a TONPS = 69 and
TONCAT = 856 at a ratio of 1 : 40 CAT : PS. At the same time, the TONCAT = 34.4. Comparing these values with the ones at 1 : 5
TONPS is minimum under these conditions with TONPS = 43 (see ratio at higher loading reveals that the system performs better
Figure 5 and Figure S4 for time traces). The performance of with higher local concentrations within the membrane.
photosensitizer was better at lower CAT:PS ratio with the best The effect of ascorbic acid concentration is depicted in
TONPS = 78 � 1 at 1 : 5 ratio with a liposome composition of Figure 6 and Figure S5. We observed the best H2 production
100 : 1 : 5 : 1 of DMPC:(14 : 0 PEG2000 PE) : RuC9:[Mo3S13]2 . It is to between 70 mM and 150 mM. Reducing electron donor concen-
mention the irregular performance at 1 : 10 ratio, slightly differ- tration below 70 mM reduces the overall H2 evolution in a close
ent from the general linear trend, we recorded for all the other to linear relationship.
case studies. We hypothesize that further increasing of CAT:PS To check the role of electron donor lactic acid at pH 6,
ratios would yield better performance of the photosensitizer, triethylamine at pH 8 or triethanolamine at pH 8 were tested.
which is the only component in the system that contains a
precious element. Unfortunately, increasing catalyst loading
was impeded by its precipitation during sample preparation at

Figure 6. Effect of different ascorbic acid concentrations on the H2 evolution


Figure 5. Effect of different CAT:PS ratios on the TONs results under under irradiation with 460 nm LED for 24 h of bulk liposome samples. H2
irradiation with 460 nm LED for 24 h for bulk liposome samples. H2 quantification was performed using head-space GC sampling in the gas
quantification was performed using head-space GC sampling in the gas phase. Experimental conditions: Liposome samples in Ar atmosphere with a
phase. Experimental conditions: liposome samples in Ar atmosphere with a composition of 100 : 1 : 5 : 1 of DMPC:(14 : 0 PEG2000 PE) : RuC9 : [Mo3S13]2 ,
composition of 100 : 1 : 5 : X of DMPC:(14 : 0 PEG2000 PE):RuC9:[Mo3S13]2 , with with c(DMPC) = 5 mM, c([Mo3S13]2 ) = 50 μM and c(ascorbic acid) = 150, 100,
c(DMPC) = 5 mM, X = 1, 0.5, 0.25 or 0.125 and c(ascorbic acid) = 100 mM, 70, 40 or 10 mM, pH = 6. All values were acquired in triplicate with error bars
pH = 6. presenting the standard deviation.

Chem. Eur. J. 2023, 29, e202302284 (5 of 11) © 2023 The Authors. Chemistry - A European Journal published by Wiley-VCH GmbH
15213765, 2023, 72, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/chem.202302284 by Indian Institute of Technology Indore, Wiley Online Library on [13/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Research Article
Chemistry—A European Journal doi.org/10.1002/chem.202302284

The pH was chosen based on the best electron donating


properties of the respective electron donor.[33] However, only
the beforementioned ascorbate electron donor yielded H2
evolution after 24 hours of irradiation (see Figure S7), which
highlights that ascorbic acid is the best suitable electron donor
for H2 evolution with RuC9 and : [Mo3S13]2 under the inves-
tigated conditions in liposomes.
We also performed in situ microelectrochemical studies to
monitor the H2 evolution using SECM in generation/collection
mode.[39] The advantage of in situ SECM measurements com-
pared to head space GC is the possibility to measure
continuously the H2 evolution without requiring a sampling
step. Hence, these in situ SECM experiments on light-active
DMPC liposomes gave insights about the first 60 minutes of
light-driven H2 evolution. The Pd microsensors[26] were posi-
tioned at a distance of approx. 50 μm to the surface of a track-
etched membrane, which separated the lower compartment
containing either the light-active DMPC-based liposomes or the
catalyst [Mo3S13]2 and RuC9 in solution without lipids. (Exper-
imental details, see Supporting Information). A potential of
600 mV vs. quasi Ag/AgCl reference electrode (QRE) was
applied and the current resulting from the palladium-hydride
(Pd H) formation[26] was recorded to determine the dissolved H2
evolution, diffusing from the lower compartment to the upper
compartment, using Faraday’s law (integration of the current
response). Figure 7A shows exemplary current response curves
with and without illumination. A large capacitive current and
associated long decay times are observed, which are related to
the large electroactive area of the Pd-modified, etched Au Figure 7. A) Amperometric i-t curves, recorded via in situ SECM experiments.
Liposome samples with a composition of 100 : 1 : 5 : 1 of DMPC : (14 : 0
microelectrodes. The immediate increase in anodic current at PEG2000 PE) : RuC9 : [Mo3S13]2 , and samples where phospholipids were
the start of the illumination is related to the photocurrent of absent. Illumination was from the bottom of the transparent electrochemical
RuC9.[40] As shown in Figure 7B, a H2 evolution rate of 3.5 � cell using a 400 μm optical fiber connected to a 21.8 mW blue LED
(λ = 470 nm). Vertical dashed lines indicate the start of illumination. B) Bar
0.31 fmol/s (n = 3) was obtained for the light-active active diagram of H2 evolution reaction rate, obtained from in situ SECM irradiation
DMPC liposomes. In contrast, the SECM measurements with experiments of samples with and without phosopholipids during the first 60
[Mo3S13]2 and RuC9 in the absence of phospholipids in the minutes.
lower compartment resulted in 0.08 � 0.02 fmol/s under the
same experimental conditions. These results show that under
the given experimental conditions, significant H2 evolution peroxide (H2O2) measurements at porcine blood cells using a
activity is only achieved with light-active DMPC based vesicles similar set-up with the two-compartment cell, detection of H2O2
and not from the individual molecular components in aqueous started instantly after adding the porcine blood cells to the
solution. Comparing the SECM results with bulk catalysis lower compartment.[42] Therefore, the delay in H2 evolution
(Figure 2A), it has to be considered that this significant lower observed in our system seems to be related either to some
rate observed in SECM is related to the different experimental delay in the photocatalysis or in low rates in the first minutes of
setup. The solution in the lower compartment was not stirred, irradiation so that the sensitivity of the measurements is not
also, the photon flux, respectively the light intensity is different sufficient to detect this initial low concentrations.[26] Hence, for
for the two approaches, clearly affecting the H2 production rate. the SECM measurements, HER catalysis can be observed after
Moreover, as shown in the setup (Figure 2B), only an area of approx. 26 minutes, with an average reaction rate of 3.5 fmol/s.
approx. 0.004 mm2 and a volume of 0.0002 mm3 is probed. Previous studies on light-driven reactions with liposome
Nonetheless, the continuous measurements allowed us to reported strong effects of the surface charge and the diffusion
monitor the H2 evolution over the first 60 minutes of the mobility within the membrane on the photocatalytic
experiment, and interestingly, the onset for both sets of conversion.[8,14,43,44] We therefore screened various lipid composi-
experiments (with and without liposomes) is delayed by ca. 21 tions to find the optimum composition. As mentioned above,
to 35 minutes with an average delay of approx. 26 min from the DMPC has zwitterionic head groups and a transition temper-
start of irradiation (see Figure 7A for example of amperometric ature at around room temperature. Exchanging the main lipid
i-t curves). In contrast, in other in situ SECM H2 measurements to the zwitterionic fluid phase DOPC, zwitterionic gel phase
with heterogenized cobaloxime catalysts, H2 evolution started DPPC, or negatively charged fluid DOPG reduced the H2
directly upon illumination.[41] Also in SECM-based hydrogen evolution (after 24 h irradiation) to TONDOPC = 9 � 3, TONDPPC =

Chem. Eur. J. 2023, 29, e202302284 (6 of 11) © 2023 The Authors. Chemistry - A European Journal published by Wiley-VCH GmbH
15213765, 2023, 72, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/chem.202302284 by Indian Institute of Technology Indore, Wiley Online Library on [13/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Research Article
Chemistry—A European Journal doi.org/10.1002/chem.202302284

11 � 4 and TONDOPG = 10 � 3, which is comparable to the results Excited state electron transfer dynamics
in absence of any phospholipid (TONCAT = 18 � 3) (Figure 8).
These results are surprising, as in previous reports with similar The effect of the lipid bilayer matrix on excited state electron
amphiphilic ruthenium-based photosensitizers it was shown transfer dynamics was investigated using steady state and time
that higher mobility within the lipid bilayer membrane are resolved emission quenching studies. In principle, it is possible
advantageous for light-driven reduction reactions as this is that upon PS excitation, the initial electron transfer takes place
advantageous for intermolecular electron transfer dynamics either i) from the sacrificial electron donor to the excited PS
within the soft matter matrix.[14,43] The opposite effect was (reductive quenching), or ii) from the excited PS to the catalyst
reported for a different type of photosensitizer based on oleic (oxidative quenching). In a so called Stern-Volmer experiment,
acid capped quantum dots, where the advantageous effect of a the excited state reaction dynamics of the photosensitizer were
rigid lipid bilayer is to protect the photosensitizer and yield resolved and the following experiments were conducted: life-
more hydrogen evolution.[14,43,44] In our case, no such trend is time and emission intensity of i) liposomes with constant
observed and the membrane environments with mobile fluid amount of RuC9 PS (no catalyst) and various amounts of
aggregation state (DOPC, DOPG) perform equally poor as the sacrificial electron donor ascorbate, ii) liposomes with constant
membrane with a rigid gel-phased aggregation state (DPPC). amount of RuC9 PS and varying amount of [Mo3S13]2 catalyst in
Only the membrane which operates at the transition temper- the membrane, iii) liposomes with constant amount of RuC9 PS
ature, DMPC, yields significant hydrogen production. and varying amount of [Mo3S13]2 catalyst in the membrane at a
constant ascorbate donor concentration. A typical dataset is
shown in Figure 9A–B and in the Supporting Information on
page 7–12, respectively. For the Stern-Volmer analysis, the
I t
intensities and lifetimes were then plotted as I0 and t0 vs.
quencher concentration [Q] (quencher = catalyst or electron
donor) where I0 and I are emission intensities in the absence
and presence of quencher; τ0 and τ are excited state lifetimes in
the absence and presence of quencher, respectively.

I0
¼ 1þKsv ½Q�¼ 1þkq t0 ½Q� (1)
I

Applying the Stern-Volmer equation [Equation (1)] as a


linear regression yields the Stern-Volmer constant Ksv and the
quenching constant kq. From the slope of the linear regression
I
of the I0 vs. [Q] plot, Ksv and the quenching constant kq [see
Equation (1)] can be derived.
Experimentally, it was found that the excited state lifetime
τ0 of RuC9 always ranges around 400 ns (see Table 3) and
Figure 8. Effect of the main phospholipid on the TONCAT results under remained unchanged irrespective of the quencher concentra-
irradiation with 460 nm LED for 24 h for bulk liposome samples with H2
quantification in the gas phase. Experimental conditions: Liposome samples tion. On the other hand, the emission intensity was reduced in
in Ar atmosphere with a composition of 100 : 1 : 5 : 1 of main lipid:(14 : 0 almost all cases, which lead to the conclusion that static
PEG2000 PE) : RuC9 : [Mo3S13]2 , with c(main lipid) = 5 mM, c([Mo3S13]2 ) - quenching takes place in almost all cases, where quenching
= 50 μM, c(ascorbic acid) = 100 mM (pH 6). All values were acquired in
triplicates with error bars presenting the standard deviation. was observed (see Figure 9 for exemplary data set for DMPC

Figure 9. A) Normalized emission intensity, B) lifetime measurements and C) Stern-Volmer plot as a function of catalyst concentration from DMPC liposome
samples with a composition of 100 : 1 : 5:X of DMPC:(14 : 0 PEG2000 PE) : RuC9 : [Mo3S13]2; c(main lipid) = 5 mM at various loadings of [Mo3S13]2 prepared with
10 mM phosphate buffer in presence of 100 mM ascorbic acid (pH 6) as sacrificial electron donor.

Chem. Eur. J. 2023, 29, e202302284 (7 of 11) © 2023 The Authors. Chemistry - A European Journal published by Wiley-VCH GmbH
15213765, 2023, 72, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/chem.202302284 by Indian Institute of Technology Indore, Wiley Online Library on [13/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Research Article
Chemistry—A European Journal doi.org/10.1002/chem.202302284

Table 3. Summary of Stern-Volmer constants, quenching rate constants and lifetime from mechanistic studies. Experimental conditions: DMPC, DPPC, DOPC
and DOPG liposome samples with a composition of 100 : 1 : 5 of main lipid:(14:0 PEG2000 PE):RuC9, and c(main lipid) = 5 mM prepared in 10 mM phosphate
buffer containing sodium ascorbate as sacrificial electron donor; and DMPC, DPPC, DOPC and DOPG liposome samples with a composition of 100 : 1 : 5:X of
DMPC : (14 : 0 PEG2000 PE) : RuC9 : [Mo3S13]2 , with c(main lipid) = 5 mM at various loadings of [Mo3S13]2 , prepared in 10 mM phosphate buffer with or without
100 mM ascorbic acid (pH 6) as sacrificial electron donor. Values are reported as mean values + / standard deviation.
Samples Lipid Ksv (L moL 1) kq (L moL 1
s 1) τ0 (ns) Quenching mechanism
7
Varying donor concentration, no catalyst DMPC 32 � 3 (8 � 2) · 10 390 static
DPPC 23 � 2 (5 � 1) · 107 466 static
7
DOPC 4�1 (1.0 � 0.3) · 10 389 static
DOPG 0.8 � 0.1 (1.7 � 0.3) · 106 455 negligible quenching

Varying catalyst concentration, no donor DMPC (3.6 � 0.3) · 104 (9 � 1) · 1010 383 static
4
DPPC (2.0 � 0.3) · 10 (5 � 1) · 1010 415 static
DOPC (9 � 1) · 103 (1.9 � 0.4) · 1010 486 static
3 9
DOPG (2.9 � 0.6) · 10 (6 � 2) · 10 498 static

Varying catalyst concentration, constant donor concentration DMPC (3.0 � 0.6) · 104 (8 � 2) · 1010 390 static
DPPC (1.5 � 0.2) · 103 (4 � 1) · 109 342 static
3 10
DOPC (5 � 1) · 10 (1.4 � 0.8) · 10 372 static
DOPG (2.0 � 0.3) · 103 (4 � 1) · 109 485 static

and Supporting Information on page 7–10).[45] A static quench-


ing mechanism is typical for excited state charge transfer where
the photosensitizer and the quencher form an encounter
complex prior to photoexcitation. All molecules in the encoun-
ter complex are quenched completely, while the remaining
photosensitizer molecules can luminesce with their normal
luminescence lifetime. In fact, static quenching seems to be the
common mechanism with positively charged amphiphilic
ruthenium trisbipyridine type photosensitizers within a lipid
bilayer membrane and a water soluble, negatively charged
quencher.[5,46] It is therefore concluded that a high local
concentration of ascorbate is attracted to the liposome
membranes loaded with cationic photosensitizer. In fact, the
only case with only negligible emission quenching, hence no Figure 10. Comparison of normalized emission intensities from DMPC lip-
osome samples with c(main lipid) = 5 mM, prepared in 10 mM phosphate
initial light-driven electron transfer, was the negatively charged buffer with or without 50 μM HER catalyst and 100 mM ascorbic acid (pH 6)
lipid bilayer of DOPG in presence of negatively charged as sacrificial electron donor (ED).
ascorbate electron donor (see Supporting Information on
page 7–10), which is due to electrostatic repulsion and is in line
with the poor performance in light-driven hydrogen evolution conclusion that excited state electron transfer takes place via
in DOPG vesicles. Comparing the other lipids with DMPC and both reductive and oxidative quenching with the electron
DOPG, the quenching by ascorbate ions ranges is either one donor and catalyst (steps 2 in Figure 1B).
third less effective (DPPC, gel phase) or approximately factor 10 To analyze the competition of oxidative and reductive
less effective (DOPC, fluid phase) than the best performing host quenching, further Stern-Volmer experiments were performed
lipid DMPC (transition temperature between gel and fluid using different CAT:PS ratios in the presence of a constant
phase). concentration of electron donor. They showed that oxidative
Considering the quenching of excited state photosensitizer quenching by the catalyst still took place with similar KSV and
by the catalyst, the trend is the same: Quenching in DMPC is quenching constants kq compared to conditions where no
more effective than in all other lipid matrices, in the series competing reductive quencher was present (see Table 3),
DMPC > DPPC > DOPC > DOPG. The quenching constant is sig- including the fact that quenching is fastest in DMPC (kq = (8 �
nificantly higher than for quenching by the electron donor. 2) · 1010 L moL 1 s 1).
However, at catalytic conditions the overall quenching by The only exception is the gel-phase lipid DPPC with kq =
ascorbate and catalyst is similar and the combination leads to (4 � 1) · 109 L moL 1 s 1 which showed factor 10 less effective
even stronger quenching (see Figure 10), leading to the quenching by the catalyst in presence of ascorbate compared

Chem. Eur. J. 2023, 29, e202302284 (8 of 11) © 2023 The Authors. Chemistry - A European Journal published by Wiley-VCH GmbH
15213765, 2023, 72, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/chem.202302284 by Indian Institute of Technology Indore, Wiley Online Library on [13/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Research Article
Chemistry—A European Journal doi.org/10.1002/chem.202302284

to DPPC liposomes in absence of the electron donor, making sensors allowed us to continuously measure the H2 evolution
oxidative quenching in DPPC one order of magnitude slower over the first 60 minute of irradiation and revealed an onset of
compared to DMPC. The fact that the presence of both measurable H2 evolution after around 26 min of the DMPC
quenchers makes the oxidative quenching by the catalyst less based photoactive DMPC liposomes, which indicates a relatively
favorable in DPPC, hints towards a competition of catalyst and slow activation mechanism, probably due to charge recombina-
electron donor in the gel phase lipid bilayer of DPPC which can tion processes. Future studies on the lipid – molecule
be explained with the low mobility of RuC9 and [Mo3S13]2 at interactions will be performed to fully understand and predict
the membrane-water interface. It also allows to conclude that, the performance of any new light-driven system on lipid
although oxidative and reductive quenching take place at the bilayers.
same time, oxidative quenching initiates the main productive
reaction pathway through initial electron transfer to the catalyst
in DPPC liposomes. Experimental Section
In summary, DMPC performs best in the catalysis experi-
ments, and the excited state PS is most efficiently quenched by Materials
the CAT and the electron donor, suggesting that initial
RuC9 and Na2[Mo3S13] x 5H2O were synthesized as previously
quenching efficiency determines the light-driven H2 evolution reported in our studies.[1–3,4,5]
of [Mo3S13]2 -CAT in lipid bilayers. DMPC is the lipid which has a
Lipids 1,2-dimyristoyl-sn-glycero-3-phosphocholine 14:0 PC (DMPC),
transition temperature at room temperature and therefore a
1,2-dimyristoyl-sn-glycero-3-phosphoethanolamine-N-[meth-
mix of fluid and gel phase regions (see Figure 11 for visual- oxy(polyethylene glycol)-2000] (14:0 PEG2000 PE), 1,2-dioleoyl-sn-
ization) which seems to provide the best local environment for glycero-3-phosphocholine 18 : 1 (Δ9-Cis) PC (DOPC), 1,2-dioleoyl-sn-
charge transfer between CAT, PS and electron donor. glycero-3-phospho-(1’-rac-glycerol) 18 : 1 (Δ9-Cis) (DOPG) and 1,2-
dipalmitoyl-sn-glycero-3-phosphocholine 16 : 0 PC (DPPC), were
purchased from Avanti Polar Lipids, L-ascorbic Acid 99 % was
purchased from Sigma Aldrich, methanol, acetonitrile, chloroform,
Conclusions lactic acid 99 %, triethylamine and triethanolamine with analytical
grade of purity were purchased from Sigma Aldrich.
The present study demonstrates that material integration of the
For Pd-microsensor fabrication and experiments di-potassium
molecular photosensitizer RuC9 and the molybdenum sulfide
hydrogen phosphate were purchased from VWR Chemicals, Darm-
HER catalyst [Mo3S13]2 into biomimetic lipid bilayers produces stadt, Germany. 1,1’-ferrocendimethanol (97 %) was purchased from
stable vesicle structures and can lead to significant improve- Acros Organics, Fair Lawn, USA. Potassium tetrachloropalladate(II)
ments of catalytic performance in water when choosing the (99.99 %) was purchased from Alfa Aesar, Thermo Fisher GmbH,
optimum lipid environment. Zwitterionic DMPC lipid bilayers Kandel, Germany. Potassium chloride, sodium chloride and potas-
with a transition temperature at the experimental conditions sium dihydrogen phosphate were purchased from Merck KGaA,
Darmstadt, Germany. The SECM studies were performed in a three-
provide the optimal environment for this specific photosensi-
electrode setup with the Pd-modified microelectrode as WE, a Ag/
tizer-catalyst combination, by promoting effective charge trans- AgCl quasi reference electrode (QRE) and a Pt wire as CE using a
fer between the PS, the electron donor and the catalyst. Lipids custom-built SECM two-compartment electrochemical cell. All
with higher or lower transition temperatures (DOPC, DPPC, experiments were controlled via a CH Instrument potentiostat
DOPG) seem to block the activity in light-driven HER and (Modell: 760E CH Instruments, Texas, USA). The two compartments
(lower and upper compartment) of the electrochemical cell were
perform equally poor as the reaction in purely aqueous environ-
separated by a track-etched polycarbonate membrane with a pore
ment. These findings were underlined by quenching studies, density of about 4.5 · 1012 pores/m2 and a pore size of 50 nm to
revealing that initial photoexcitation of the photosensitizer is prevent fouling of the Pd-modified microelectrode by the lip-
followed by simultaneous reductive and oxidative quenching osomes. Illumination occurred from the bottom of the transparent
which are fastest in DMPC compared to the other lipid electrochemical cell using a 400 μm diameter optical fiber (MT-
environments. The DMPC membrane is at the transition temper- 28L01, Thorlabs GmbH, Bergkirchen, Germany) connected to a
21.8 mW blue LED (λ = 470 nm, M470F3, Thorlabs GmbH).
ature under experimental conditions and shows preferential
electron transfer dynamics for the excited state PS which might TEM analysis were carried out with a Jeol 1400 TEM via negative
be due to the presence of a mixture of gel phase and fluid staining technique using uranyl acetate as suitable electron-dense
phase domains. In situ SECM measurements using Pd micro- material to provide high contrast.
Particle size measurements were performed with a Zetasizer Pro
from Malvern Panalytical by Dynamic Light Scattering (DLS)-Non
Invasive Back Scatter (NIBS) using 900 μL of sample in a disposable
polystyrene cell (DTS0012) with a measurement angle of 173°, 13°
in a diameter range of 0.3 nm–10 μm at 20 °C. Zeta potential
measurements were performed with a Zetasizer Pro from Malvern
Panalytical by Mixed-Mode Measurement phase analysis light
scattering (M3-PALS) using a disposable folded capillary cell
Figure 11. Representation of the mix of fluid and gel phase for DMPC, (DTS1070) in a diameter range of 3.8 nm–100 μm.
compared to the pure gel phase (DPPC Tm = 41 °C) and gel phase (DOPG,
DOPC Tm = 18 °C and 17 °C respectively), at the temperature applied
during light-driven catalysis experiments and mechanistic studies.

Chem. Eur. J. 2023, 29, e202302284 (9 of 11) © 2023 The Authors. Chemistry - A European Journal published by Wiley-VCH GmbH
15213765, 2023, 72, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/chem.202302284 by Indian Institute of Technology Indore, Wiley Online Library on [13/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Research Article
Chemistry—A European Journal doi.org/10.1002/chem.202302284

H2 evolution was quantified using a Bruker Scion SQ GC/MS by Conflict of Interests


injecting manually 100 μL of the head space. Sample preparation
under inert conditions was performed in an MBRAUN LABSTAR There are no conflicts to declare.
glovebox and pH values were measured with a Mettler Toledo
FiveGoTM pH meter. Photocatalysis was performed in a 3D printed
photoreactor from project C6 in Catalight network equipped with a
LZ1-00DB00 Dental Blue LED emitter (λ = 460 nm). UV-Vis Absorp- Data Availability Statement
tion spectroscopy was performed on a V-760 JASCO UV-VIS-NIR
Spectrophotometer. [Gas-tight] Quartz glass cuvettes (d = 10.0 mm) The data that support the findings of this study are available
were used for optical measurements. Emission intensity and lifetime
from the corresponding author upon reasonable request.
measurements were performed, respectively, using a Horiba Jobin-
Yvon FluoroMax Plus C automated benchtop spectrofluorometer
equipped with a 150 W Xe arc excitation lamp (horizontal, Keywords: catalysis · electron transfer · hydrogen · liposomes ·
continuous wave), a R13456 photon-counting PMT detector (190- photochemistry
930 nm) and DeltaPro from Horiba Scientific using a 451 nm pulsed
Laser source (Class 3B Laser Product, < 0.5 W peak in pulsed and
CW mode.
[1] N. S. Lewis, D. G. Nocera, Proc. Nat. Acad. Sci. 2006, 103, 15729–15735.
[2] R. Eisenberg, D. G. Nocera, Inorg. Chem. 2005, 44, 6799–6801.
Vesicle preparation [3] T. R. Cook, D. K. Dogutan, S. Y. Reece, Y. Surendranath, T. S. Teets, D. G.
Nocera, Chem. Rev. 2010, 110, 6474–6502.
Stocks solutions are reported in the Supporting Information. [4] F. Wang, W.-G. Wang, H.-Y. Wang, G. Si, C.-H. Tung, L.-Z. Wu, ACS Catal.
Vesicles were prepared as previously reported,[5,28] details can be 2012, 2, 407–416.
found in the Supporting Information. [5] S. Rodríguez-Jiménez, H. Song, E. Lam, D. Wright, A. Pannwitz, S. A.
Bonke, J. J. Baumberg, S. Bonnet, L. Hammarström, E. Reisner, J. Am.
Chem. Soc. 2022, 144, 9399–9412.
Photocatalysis [6] M. Hansen, S. Troppmann, B. König, Chem. Eur. J. 2016, 22, 58–72.
[7] H. Hu, Z. Wang, L. Cao, L. Zeng, C. Zhang, W. Lin, C. Wang, Nat. Chem.
Sample preparation was done under Argon atmosphere within a 2021, 13, 358–366.
glovebox (O2 content < 0.5 ppm) using 250 μL of liposomes [8] S. Troppmann, B. König, Chem. Eur. J. 2014, 20, 14570–14574.
solution at 5 mM main lipid concentration and different amounts of [9] R. Becker, T. Bouwens, E. C. F. Schippers, T. van Gelderen, M. Hilbers, S.
the electron donor solution, to obtain different concentrations (see Woutersen, J. N. H. Reek, Chem. Eur. J. 2019, 25, 13921–13929.
[10] B. Limburg, J. Wermink, S. S. Van Nielen, R. Kortlever, M. T. M. Koper, E.
Figure S5 and see preparation of stock solutions), which were
Bouwman, S. Bonnet, ACS Catal. 2016, 6, 5968–5977.
combined in an 8 mL WICOM WIC 41600/333 clear glass screw neck [11] N. Sinambela, J. Bösking, A. Abbas, A. Pannwitz, ChemBioChem 2021, 22,
GC vial (ND13). The vial was then sealed, transferred into a 3D 3140–3147.
printed photoreactor, and irradiated by LZ1-00DB00 Dental Blue as [12] S. Takizawa, T. Okuyama, S. Yamazaki, K. Sato, H. Masai, T. Iwai, S.
LED light source (λ = 460 nm, 800 to 1250 mW) with a current of Murata, J. Terao, J. Am. Chem. Soc. 2023, 35, 37.
0.7 A and voltage of 4 V. Simultaneously, 3 identically prepared [13] A. Pannwitz, D. M. Klein, S. Rodríguez-Jiménez, C. Casadevall, H. Song, E.
vials were irradiated while stirring for 24 hours, using two ring Reisner, L. Hammarström, S. Bonnet, Chem. Soc. Rev. 2021, 50, 4833–
4855.
holders to improve the alignment. The experiment was performed
[14] D. M. Klein, S. Rodríguez-Jiménez, M. E. Hoefnagel, A. Pannwitz, A.
at room temperature via a ventilation system in the back of the Prabhakaran, M. A. Siegler, T. E. Keyes, E. Reisner, A. M. Brouwer, S.
reactor. After 2 h, 4 h and 24 h, the evolved H2 was analyzed via Bonnet, Chem. Eur. J. 2021, 27, 17203–17212.
head-space gas chromatography from each reaction vessel. [15] R. E. P. Nau, J. Bösking, A. Pannwitz, ChemPhotoChem 2022, 6,
Turnover numbers data were calculated from the ratio between e202200158.
moles of H2 produced and moles of Na2[Mo3S13] x 5H2O or bis(2,2’- [16] M. Dave, A. Rajagopal, M. Damm-Ruttensperger, B. Schwarz, F. Nägele,
bipyridine)-(4,4’-dinonyl-2,2’-bipyridine)-ruthenium(II) (RuC9). L. Daccache, D. Fantauzzi, T. Jacob, C. Streb, Sustain. Energy Fuels 2018,
Figures of the photoreactor see Supporting Information. 2, 1020–1026.
[17] H. I. Karunadasa, E. Montalvo, Y. Sun, M. Majda, J. R. Long, C. J. Chang,
Science 2012, 335, 698–702.
[18] Z. Huang, W. Luo, L. Ma, M. Yu, X. Ren, M. He, S. Polen, K. Click, B.
Garrett, J. Lu, K. Amine, C. Hadad, W. Chen, A. Asthagiri, Y. Wu, Angew.
Acknowledgements Chem. Int. Ed. 2015, 54, 15181–15185.
[19] D. Recatalá, R. Llusar, A. L. Gushchin, E. A. Kozlova, Y. A. Laricheva, P. A.
Abramov, M. N. Sokolov, R. Gómez, T. Lana-Villarreal, ChemSusChem
A. P. acknowledges the Vector Stiftung (project number P2019- 2015, 8, 148–157.
0110) for financial support and the authors thank the Deutsche [20] M. L. Grutza, A. Rajagopal, C. Streb, P. Kurz, Sustain. Energy Fuels 2018, 2,
1893–1904.
Forschungsgemeinschaft (DFG, project TRR234 “CataLight” [21] D. Kowalczyk, P. Li, A. Abbas, J. Eichhorn, P. Buday, M. Heiland, A.
project number 364549901, projects A5, B8, C4) for financial Pannwitz, F. H. Schacher, W. Weigand, C. Streb, D. Ziegenbalg,
support. We thank Renate Kunz and Paul Walther from the ChemPhotoChem 2022, 6, e202200044.
[22] I. Romanenko, A. Rajagopal, C. Neumann, A. Turchanin, C. Streb, F. H.
Central Facility for Electron Microscopy, Ulm University, for help Schacher, J. Mater. Chem. A 2020, 8, 6238–6244.
with Transmission Electron Microscopy measurements. Some [23] A. Nabiyan, J. B. Max, C. Neumann, M. Heiland, A. Turchanin, C. Streb,
figures were created with BioRender.com as indicated in the F. H. Schacher, Chem. Eur. J. 2021, DOI 10.1002/chem.202100091.
[24] A. Rajagopal, E. Akbarzadeh, C. Li, D. Mitoraj, I. Krivtsov, C. Adler, T.
figure captions. Open Access funding enabled and organized by Diemant, J. Biskupek, U. Kaiser, C. Im, M. Heiland, T. Jacob, C. Streb, B.
Projekt DEAL. Dietzek, R. Beranek, Sustain. Energy Fuels 2020, 4, 6085–6095.
[25] J. Eichhorn, P. Hofmann, B. Bagemihl, C. Streb, S. Rau, F. H. Schacher, J.
Mater. Chem. A 2023, 11, 11334–11340.
[26] J. Kund, J. Romer, E. Oswald, A.-L. Gaus, M. Küllmer, A. Turchanin, M.
Delius, C. Kranz, ChemElectroChem 2022, 9, 2–7.

Chem. Eur. J. 2023, 29, e202302284 (10 of 11) © 2023 The Authors. Chemistry - A European Journal published by Wiley-VCH GmbH
15213765, 2023, 72, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/chem.202302284 by Indian Institute of Technology Indore, Wiley Online Library on [13/01/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Research Article
Chemistry—A European Journal doi.org/10.1002/chem.202302284

[27] A. Müller, V. Wittneben, E. Krickemeyer, H. Bögge, M. Lemke, Z. Anorg. [39] J. L. Fernández, C. G. Zoski, J. Phys. Chem. C 2018, 122, 71–82.
Allg. Chem. 1991, 605, 175–188. [40] K. Jiang, H. Xie, W. Zhan, Langmuir 2009, 25, 11129–11136.
[28] A. Pannwitz, H. Saaring, N. Beztsinna, X. Li, M. A. Siegler, S. Bonnet, [41] E. Oswald, A.-L. Gaus, J. Kund, M. Küllmer, J. Romer, S. Weizenegger, T.
Chem. Eur. J. 2021, 27, 3013–3018. Ullrich, A. K. Mengele, L. Petermann, R. Leiter, P. R. Unwin, U. Kaiser, S.
[29] E. Rideau, R. Dimova, P. Schwille, F. R. Wurm, K. Landfester, Chem. Soc. Rau, A. Kahnt, A. Turchanin, M. Delius, C. Kranz, Chem. Eur. J. 2021, 27,
Rev. 2018, 47, 8572–8610. 16896–16903.
[30] A.-L. Robson, P. C. Dastoor, J. Flynn, W. Palmer, A. Martin, D. W. Smith, A. [42] P. F. F. Almeida, W. L. C. Vaz, T. E. Thompson, Biochemistry 1992, 31,
Woldu, S. Hua, Front. Pharmacol. 2018, 9, 80. 6739–6747.
[31] U. Baxa, Methods Mol. Biol. 2018, 1682, 73–88. [43] S. Troppmann, E. Brandes, H. Motschmann, F. Li, M. Wang, L. Sun, B.
[32] B. Ruozi, D. Belletti, A. Tombesi, G. Tosi, L. Bondioli, F. Forni, M. A. König, Eur. J. Inorg. Chem. 2016, 2016, 554–560.
Vandelli, Int. J. Nanomed. 2011, 6, 557. [44] S. Troppmann, B. König, ChemistrySelect 2016, 1, 1405–1409.
[33] Y. Pellegrin, F. Odobel, Comptes Rendus Chim. 2017, 20, 283–295. [45] V. Balzani, P. Ceroni, A. Juris, Photochemistry and Photophysics. Concepts,
[34] B. Limburg, E. Bouwman, S. Bonnet, ACS Catal. 2016, 6, 5273–5284. Research, Applications, Wiley-VCH, Weinheim, 2014.
[35] J. L. Grant, K. Goswami, L. O. Spreer, J. W. Otvos, M. Calvin, J. Chem. Soc. [46] H. Song, A. Amati, A. Pannwitz, S. Bonnet, L. Hammarström, J. Am.
Dalton Trans. 1987, 2105. Chem. Soc. 2022, 144, 19353–19364.
[36] J. Hawecker, J. M. Lehn, R. Ziessel, J. Chem. Soc. Chem. Commun. 1985,
56–58.
[37] A. Nakada, K. Koike, T. Nakashima, T. Morimoto, O. Ishitani, Inorg. Chem.
2015, 54, 1800–1807. Manuscript received: July 17, 2023
[38] J. Li, C. A. Triana, W. Wan, D. P. Adiyeri Saseendran, Y. Zhao, S. E. Balaghi, Accepted manuscript online: September 12, 2023
S. Heidari, G. R. Patzke, Chem. Soc. Rev. 2021, 50, 2444–2485. Version of record online: November 8, 2023

Chem. Eur. J. 2023, 29, e202302284 (11 of 11) © 2023 The Authors. Chemistry - A European Journal published by Wiley-VCH GmbH

You might also like