Play all audios:
ABSTRACT Using a magnetron sputtering approach that allows size-controlled formation of nanoclusters, we have created palladium nanoclusters that combine the features of both heterogeneous
and homogeneous catalysts. Here we report the atomic structures and electronic environments of a series of metal nanoclusters in ionic liquids at different stages of formation, leading to
the discovery of Pd nanoclusters with a core of _ca_. 2 nm surrounded by a diffuse dynamic shell of atoms in [C4C1Im][NTf2]. Comparison of the catalytic activity of Pd nanoclusters in alkene
cyclopropanation reveals that the atomically dynamic surface is critically important, increasing the activity by a factor of _ca_. 2 when compared to compact nanoclusters of similar size.
Catalyst poisoning tests using mercury and dibenzo[_a,e_]cyclooctene show that dynamic Pd nanoclusters maintain their catalytic activity, which demonstrate their combined features of
homogeneous and heterogeneous catalysts within the same material. Additionally, kinetic studies of cyclopropanation of alkenes mediated by the dynamic Pd nanoclusters reveal an observed
catalyst order of 1, underpinning the pseudo-homogeneous character of the dynamic Pd nanoclusters. SIMILAR CONTENT BEING VIEWED BY OTHERS SITE-SPECIFIC ELECTRODEPOSITION ENABLES
SELF-TERMINATING GROWTH OF ATOMICALLY DISPERSED METAL CATALYSTS Article Open access 11 September 2020 DYNAMIC RESTRUCTURING OF SUPPORTED METAL NANOPARTICLES AND ITS IMPLICATIONS FOR
STRUCTURE INSENSITIVE CATALYSIS Article Open access 07 December 2021 SULFUR STABILIZING METAL NANOCLUSTERS ON CARBON AT HIGH TEMPERATURES Article Open access 25 May 2021 INTRODUCTION
Catalysts are used in nearly 80 % of industrial processes1, demanding the maximum catalytic efficiency of rare elements, such as Au, Pt and Pd2,3,4. Increasing the metal active surface area
by utilising catalysts in the form of nanoclusters is one of the most powerful approaches5,6,7,8,9. When the size of metal shrinks down to the nanoscale, the structure and dynamics and thus
properties of the nanocluster cannot be interpreted as a linear function of its size, which makes it challenging to correlate them with their catalytic performance7,10,11,12. In order to
establish a structure–property relationship for nanoclusters and the link with their catalytic performance, we must explore the series of related nanoclusters at their boundaries between
discrete atoms and crystalline nanoclusters, with atomic precision, and using a set of complementary analytical methods, thus shedding light on their nature13. In this context, magnetron
sputtering offers exciting new opportunities for the manufacture of novel catalytic materials with controlled sizes and shapes of metal nanoclusters on a variety of supports14,15,16,17,18.
Crucially, this approach generates highly active clean surfaces, since no chemical stabilisers and/or solvents are involved in the process16,19. Among the supports, ionic liquids (ILs) are
unique because they offer a soft, liquid environment for relatively free (in comparison to solid supports) nanoclusters to be studied in their native state. Furthermore, ILs possess low
vapour pressure allowing their use under the high vacuum conditions of magnetron sputtering20,21. However, atomic-scale analysis of metal nanoclusters in IL, for example, by transmission
electron microscopy (TEM), remains a challenge. The current sample preparation approaches using stabilisers or solvents22,23,24 may alter the environment surrounding the nanocluster, thus
potentially inducing structural changes to the nanocluster and therefore hide important information on its shape and size. To the best of our knowledge, metal nanoclusters in ILs deposited
by magnetron sputtering have not been observed with atomic resolution in their native state so far. As a result, the relationship between metal nanoclusters structure and their catalytic
activity in ILs remains largely underexplored and widely debated. Herein we determine the atomic structures and electronic environments of a series of metal nanoclusters in ILs at different
stages of formation, leading to a discovery of Pd nanoclusters with a core of ca. 2 nm surrounded by a diffuse dynamic shell of atoms in 1-butyl-3-methylimidazolium
bis(trifluoromethylsulfonyl)imide ([C4C1Im][NTf2]) that combines the features of homogeneous and heterogeneous catalysts within the same material. RESULTS AND DISCUSSION SYNTHESIS OF PD
NANOCLUSTERS The magnetron sputtering metal deposition process involves the elastic collision of argon ions with a highly pure metallic target, resulting in the expulsion of atoms or
clusters from the target that are deposited onto a support material (Fig. 1a and Supplementary Fig. 1)14. Previously, it was proposed that the growth of metal clusters occurs on the IL
surface/near surface or in the bulk IL depending on the magnetron sputtering conditions and physical properties of the employed IL25,26,27. However, a variety of parameters, including the
applied potential, argon (Ar) working pressure and work distance (distance between the metal target and the support), can influence: (i) whether metal atoms or a few metal atom cluster are
ejected from the target; (ii) the possible formation of metal clusters in the gas phase; (iii) the kinetic energy of the metal atoms/clusters landing on the IL and thus how deeply they will
penetrate. Additionally, the physical properties of the IL (i.e. surface tension) or substituent functional groups also play a major role in the metal cluster growth20. In this work,
magnetron sputtering depositions of Pd species into ILs were carried out using ultraclean conditions with a background-pressure of 5.3 × 10-6 Pa, work pressure fixed at 4.0 × 10−1 Pa using
Ar gas (99.9995%) and a high purity Pd target (99.995%). Theoretical simulations for different potentials were performed prior to Pd depositions in IL to optimise the ejection of single Pd
atoms (see Supplementary Methods for further details). The simulations showed the highest probability for single Pd atoms to be sputtered from the target at applied potentials up to 800 eV,
with two- or three-atom Pd clusters being minority species (i.e. at 320 eV, 31 single Pd atoms are ejected for each two-atom cluster—Supplementary Fig. 2). Therefore, the Pd depositions in
ILs were performed using a potential applied within the range displayed in Supplementary Fig. 2. The only parameters varied during the depositions were the Pd concentration and the nature of
IL anion. Further details of the magnetron sputtering parameters can be found in Supplementary Table 1 and Supplementary Methods. As expected, the Pd concentration in ILs increased linearly
with the Pd deposition time (Supplementary Fig. 3). Therefore, the resulting materials were named as _X_Pd@_Y_ where _X_ is Pd deposition time in min and _Y_ is the nature of the anion in
[C4C1Im]. _Y_ as the IL cation was kept constant. For instance, Pd species deposited for 5 min (Pd concentration of 0.12 wt%) in [C4C1Im][NTf2] would be denoted as 5Pd@[NTf2]. PD NANOCLUSTER
MORPHOLOGY AND ATOMIC STRUCTURE All TEM analyses were performed directly in IL to evaluate the true morphology and atomic structure of the Pd species. Bright-field TEM imaging revealed
variation of Pd nanocluster size distribution, from sub-1 nm in 1Pd@[NTf2] to 2–3 nm in all other ILs (Supplementary Figs. 4, 5, 9, 10 and 12). Atomic-scale investigation using
aberration-corrected scanning transmission electron microscopy (AC-STEM) for 5Pd@[NTf2] revealed unusual features—a Pd core surrounded by a diffuse shell of satellite Pd atoms, with the
overall nanocluster diameter of 2 ± 1.0 nm (Fig. 1b). The structure is dynamic with individual Pd atoms reshuffling over time, as revealed by time-series AC-STEM imaging, and henceforth is
referred to as a dynamic nanocluster (Fig. 1b and Supplementary Figs. 6–8). However, the further increase of the Pd concentration in [C4C1Im][NTf2] (Fig. 1c and Supplementary Fig. 11), or
the use of a different IL with the anion [PF6]− (Supplementary Fig. 13), produces a compact Pd nanocluster with significantly lower dynamics than 5Pd@[NTf2]. In the compact Pd nanoclusters,
atomic lattice planes (111) of face-centred cubic Pd are observed (Supplementary Figs. 9, 10 and 12), whereas in 5Pd@[NTf2] it appears to be amorphous (Fig. 1b and Supplementary Fig. 5). The
overall three-dimensional disorder of the dynamic Pd nanoclusters is also apparent in the irregular intensity distribution extracted from AC-STEM images (Fig. 1d and Supplementary Fig. 8b),
as compared to a more homogenous intensity pattern typical of the compact nanoclusters such as 60Pd@[NTf2] and 5Pd@[PF6] (Fig. 1e and Supplementary Figs. 11b and 13b). Matrix-assisted laser
desorption ionisation time of flight mass spectrometry clearly shows the presence of single Pd atoms in Pd@[NTf2] (Supplementary Figs. 14–20). The presence of single Pd atoms appears to be
strongly dependent on the type of IL, as changing the IL cation ([C4C1Im]+ to 1,2-dimethyl-3-butyl-imidazolium cation ([C4C1C1Im]+) or anion ([NTf2]−) to hexafluorophosphate anion ([PF6]−))
both have an impact on the availability of Pd atoms, such that no single Pd atoms are detected in [C4C1C1Im][NTf2] and [C4C1Im][PF6]. As the loading of the metal increases, the proportion of
Pd existing in the atomic form decreases, consistent with the AC-STEM observations that dynamic nanoclusters exist only in [C4C1Im][NTf2] and only at low loading of metal. PD NANOCLUSTER
ELECTRONIC ENVIRONMENT A combination of X-ray photoelectron spectroscopy (XPS), X-ray absorption spectroscopy (XAS) and Raman spectroscopy provided insight into the interactions between Pd
and [C4C1Im][NTf2]. The Pd _3d_ core XPS spectrum indicates that Pd nanoclusters in [C4C1Im][NTf2] exist in two different electronic environments: the metallic Pd peak at 335.9 eV and a peak
at 338.5 eV that we assign to the interaction between surface Pd atoms with [C4C1Im][NTf2] (Fig. 2a, Supplementary Figs. 21–25 and Supplementary Table 2). The integral intensity of metallic
Pd0 peak shows a relative decrease as the loading of Pd in [C4C1Im][NTf2] decreases (from 60Pd@[NTf2] to 10Pd@[NTf2]) (Fig. 2a). These results correlate with the observation in AC-STEM and
mass spectrometry, suggesting that surface Pd atoms interact with IL and that dynamic Pd nanoclusters exist mainly at low loadings. X-ray absorption near-edge structure (XANES) measurements
for 5Pd@[NTf2] and 30Pd@[NTf2] Pd _K_-edge show a similar adsorption edge (_E_0) to bulk Pd metal but with a higher white line (_H_w) intensity in the 30Pd@[NTf2] and a further _H_w
intensity increases for 5Pd@[NTf2]. This indicates _d_-electron depletion of Pd nanoclusters due to charge transfer to the more electronegative groups of the [C4C1Im][NTf2] IL (Fig. 2b)28,
consistent with the XPS results. Extended X-ray absorption fine structure (EXAFS) analysis of 5Pd@[NTf2] and 30Pd@[NTf2] show a peak at ca. 2.5 Å associated with Pd–Pd bond, same as in bulk
metal, as well as another peak at ca. 1.7 Å that we assigned to Pd atom interaction with [C4C1Im][NTf2] (Fig. 2c). Density functional theory (DFT) modelling confirmed that the Pd
coordination occurs via N and O atoms of the [NTf2]− anion for the Pd13[C4C1Im][NTf2], whereas for the Pd1[C4C1Im][NTf2] complex the coordination occurs via C-atoms and O-atoms from the
cation and anion, respectively (Fig. 2e). Moreover, the intensity and the peak corresponding to Pd–Pd distance decreases as the Pd loading decreases (30Pd@[NTf2] to 5Pd@[NTf2]) alongside to
a relative increase of the Pd-IL peak intensity, suggesting a higher number of Pd atoms interacting with the [C4C1Im] cation and/or [NTf2] anion at lower Pd loading in IL, in agreement with
the XPS analysis. Additionally, for 5Pd@[NTf2] EXAFS indicated slightly longer Pd–Pd bonds compared to 30Pd@[NTf2] and Pd bulk, which can be related to the dynamic Pd nanocluster behaviour
at relatively low Pd loading in IL (5Pd@[NTf2])29. Both XPS and XAS demonstrate a strong interaction between surface Pd atoms and [C4C1Im][NTf2] and with the core Pd atoms remaining in
metallic-like state. Raman spectroscopy allows probing the IL and shows the increase of _cis_- conformation of the [NTf2] anion, as the loading of Pd increases (Fig. 2d, Supplementary Fig.
27 and Supplementary Table 3). DFT modelling also shows that for [C4C1Im][NTf2] the _cis_- conformer is slightly more stable than the _trans_- conformer by 2.1 kJ mol−1, the addition of a Pd
atom, as in Pd1[C4C1Im][NTf2] and Pd13[C4C1Im][NTf2], increases the difference in stability of the conformers to 4.1 and 3.9 kJ mol−1, respectively (Fig. 2e and Supplementary Fig. 28).
CATALYTIC EVALUATION OF PD NANOCLUSTER The catalytic performance of the Pd@[NTf2] system was evaluated in a cyclopropanation of alkenes where Pd facilitates carbene transfer from ethyl
diazoacetate (EDA) to alkenes (Fig. 3a, b, for further details, see Supplementary Tables 4 and 5 and Supplementary Notes 1)30,31,32,33. All materials were found to be active in the
cyclopropanation of styrene, giving good conversions and selectivities towards the cyclopropane product. In order to compare relative catalytic conversions of different Pd nanoclusters, a
series of experiments in Pd@[NTf2] with a normalised metal loading in the reaction mixture were carried out (0.5 mol% Pd) (Fig. 3b; for further details see, Supplementary Table 4 and
Supplementary Notes 1). While selectivities were similar, a gradual increase in the conversion was observed with a decrease in the Pd metal concentration in IL (from 60Pd@[NTf2] to
10Pd@[NTf2]), with an abrupt increase for 7Pd@[NTf2], 5Pd@[NTf2] and 1Pd@[NTf2] (Fig. 3b and Supplementary Table 4), correlating with the transition from compact to dynamic Pd nanoclusters,
indicating a higher fraction of accessible Pd atoms in the latter. In order to verify the mode of catalysis by dynamic Pd nanoclusters, mercury poisoning tests were performed (Fig. 3c). It
is well established that heterogeneous catalysts, in contrast to homogenous catalysts, lose their catalytic activity in the presence of Hg due to the poisoning of their surface34. Indeed,
complete inhibition of the catalysis was observed for compact Pd nanoclusters (from 60Pd@[NTf2] to 10Pd@[NTf2]; Fig. 3c), suggesting the absence and/or very low population of dynamic Pd
nanoclusters in agreement with the AC-STEM observations. Remarkably, for the dynamic Pd nanoclusters, the catalytic activity was maintained for 1Pd@[NTf2] and only 32% decrease in yield was
observed for 5Pd@[NTf2] (Fig. 3c and Supplementary Table 6), suggesting a pseudo-homogeneous mode of catalysis by the dynamic nanoclusters. Additional poisoning tests were conducted for
higher Pd concentration in [C4C1Im][NTf2] and 5Pd@[PF6] (Supplementary Table 6), which were completely deactivated similarly to traditional heterogeneous catalysts. We have performed an
additional poisoning tests using dibenzo[_a,e_]cyclooctene (DCT), which selectively binds to single Pd sites34,35,36,37. DCT experiments showed that 5Pd@NTf2 and 30Pd@NTf2 did not lose
activity (Supplementary Table 7), confirming that dynamic Pd nanocluster are the active sites, rather than leached single Pd atoms. These results were further reinforced by filtration tests,
as both catalytic systems (5Pd@NTf2 and 30Pd@NTf2) did not show activity after the filtration (Supplementary Table 8). Additionally, recyclability experiments showed that 5Pd@NTf2 and
30Pd@NTf2 are recyclable and maintain their activity over three cycles (Supplementary Table 9). Furthermore, the N2 release was monitored during the course of the reaction to elucidate the
catalyst order (Fig. 3d, e and Supplementary Fig. 29). After pre-activation with styrene, the catalyst order was determined by the variable time normalisation analysis method introduced by
Bures and co-workers38,39,40, showing a catalyst order of 1.0 for 5Pd@[NTf2] and 0.7 for 30Pd@[NTf2] (Fig. 3d, e). These results further confirm the pseudo-homogeneous character of the
dynamic Pd nanoclusters, while the order determined for compact Pd nanoclusters is in the expected lower range for heterogeneous catalysts due to the inherently decreased amounts of atoms
accessible for catalysis. This effect relates to the condition that the reaction takes place only on the surface of the catalysts. TEM experiment where dynamic Pd nanoclusters were retrieved
from the reaction mixture directly onto a TEM grid demonstrated that the dynamic clusters remain unchanged in size and structure after the reaction (Supplementary Fig. 30). In summary, we
have explored the structure, dynamics and catalytic properties of a series of palladium nanoclusters from sub-1 to 3 nm sizes by a variety of analytical and imaging techniques, providing
atomic-scale information. Nanoclusters appear to behave as dynamic structures with properties controlled by the nature of IL (both cation and anion) and the loading of the metal in the
liquid. Anions interacting directly with palladium atoms play a particularly important role in determining the electronic state and dynamics of surface atoms. In particular, [NTf2] anions in
[C4C1Im][NTf2] have been shown to bond to surface Pd atoms and create a diffuse shell of metal atoms around the nanocluster. This phenomenon has a significant implication for the mode of
catalysis, changing the heterogeneous behaviour, exhibited by Pd nanoclusters with compact structures, to a homogeneous mode exhibited by dynamic 5Pd@[NTf2]. Our study demonstrates that the
effectiveness of nanocatalysts depends not only on the surface area of the nanoclusters but also on the dynamic behaviour of their surface atoms controlled by the support environment.
METHODS PD NANOCLUSTER PREPARATION Pd species were deposited into ILs (IL synthesis details are in the Supplementary Information) using a bespoke AJA magnetron sputtering system with a
load-lock sample transfer facility coupled to a glovebox. The Pd target (99.995%) was purchased from AJA International. IL samples were loaded through a glovebox into the magnetron
sputtering system to avoid the presence of moisture. In a typical experiment, anhydrous IL (0.750 g) is placed in a petri dish while in the glovebox. The sample was transferred to a
load-lock under a N2 environment, pumped-down to a background pressure of 5.3 × 10-5 Pa for (ca. 1 h) and then transferred to the main chamber, which reached a background pressure of 5.3 ×
10-6 Pa in ca. 1 h. The power and working pressure used in all depositions were 100 W and 4.0 × 10−1 Pa (Argon—99.9999%), respectively. SAMPLE CHARACTERISATION Pd nanoclusters deposited by
magnetron sputtering in ILs were primarily characterised by TEM, with each being analysed from at least two different deposition batches. All microscopic analyses were performed in the
absence of solvents in order to evaluate the true morphology of the Pd species deposited in the ILs. TEM measurements were performed using a JEOL 2100F FEG-TEM operated with an accelerating
voltage of 200 kV. AC-STEM measurements were performed using a JEOL 2100F scanning transmission electron microscope with a CEOS aberration corrector operated with an accelerating voltage of
200 kV. The samples were prepared in the following manner: copper mesh, holey carbon film TEM grids (Agar Scientific, UK) were glow discharged (Agar Turbo Coater, 0.2 mbar, 5 mA, 10 s)
before the addition of Pd@IL (10 µL). The Pd@IL suspension was kept on the grid for 2 min before the excess was removed using filter paper. This approach allowed the formation of ‘pools’ of
IL within the holes of the holey carbon, providing an effective contrast-free region to observe the palladium structures. XPS measurements were performed using a Kratos AXIS Ultra DLD
instrument. The chamber pressure during the measurements was 6.7 × 10−7 Pa. Wide energy range survey scans were collected at pass energy of 80 eV in hybrid slot lens mode and a step size of
0.5 eV, for 20 min. High-resolution data on the Pd 3d, C 1s, N 1s, O 1s, S 2p and F 1s photoelectron peaks were collected at a pass energy of 20 eV over energy ranges suitable for each peak
and collection times of 5 min, step sizes of 0.1 eV. The charge neutraliser filament was used to prevent the sample charging over the irradiated area. The X-ray source was a monochromated Al
Kα emission, run at 10 mA and 12 kV (120 W). The energy range for each ‘pass energy’ was calibrated using the Kratos Cu 2p3/2, Ag 3d5/2 and Au 4f7/2 three-point calibration method. The
transmission function was calibrated using a clean gold sample method for all lens modes and the Kratos transmission generator software within Vision II. The data were processed with CASAXPS
(Version 2.3.17). The high-resolution data were charge corrected to the reference F 1s signal at 688.9 eV. XAS measurements of Pd _K_-edge were performed at room temperature (B18 beamline)
at the Diamond Synchrotron Light. XANES and EXAFS spectra of a Pd foil and PdO standards were measured and the energy calibrated by aligning the respective absorption edges. The data were
calibrated and normalised by a linear pre-edge subtraction using the ATHENA software. DFT calculations were performed using version 4.1.0 of the Orca Program package41. Micro Raman
spectroscopy was performed using a Horiba Jobin Yvon LabRAM HR Raman spectrometer. Spectra were acquired using a 785 nm laser (at 24 mW power), a ×100 objective lens and a 300 µm confocal
pinhole. To simultaneously scan a range of Raman shifts and control the spectral resolution, either a 600 or 1800 lines mm−1 rotatable diffraction grating along a path length of 800 mm was
employed. Spectra were acquired using a Synapse CCD detector (1024 pixels) thermoelectrically cooled to −60 °C. CATALYTIC TESTS All reactions were prepared in a glovebox under inert
atmosphere using a 50 mL Schlenk flask equipped with a magnetic stir bar. EDA (1 mmol) was added into a solution containing the Pd catalyst (0.5 mol% Pd) and styrene (5 mmol) in CH2Cl2 (5/10
mL). The Schlenk flask was sealed with a septum cap and the reaction mixture was removed from the glovebox. After 24 h of stirring at RT, the mixture was filtered through silica gel and
volatiles were removed in vacuo. Products have been previously described and their identification was straightforward from comparison with the reported data42. Substrate conversion and
selectivity and yield of cyclopropane product were determined by 1H nuclear magnetic resonance spectroscopy using 1,2-dibromoethane (0.25 mmol) as an internal standard. DATA AVAILABILITY All
experimental and simulation data in the main text and Supplementary Materials are available upon request to the authors. CODE AVAILABILITY Molecular Dynamics codes used in the simulation
are available upon request to [email protected]. REFERENCES * Wang, A., Li, J. & Zhang, T. Heterogeneous single-atom catalysis. _Nat. Rev. Chem._ 2, 65–81 (2018). Article ADS
CAS Google Scholar * Zhang, Z. et al. Thermally stable single atom Pt/m-Al2O3 for selective hydrogenation and CO oxidation. _Nat. Commun._ 8, 16100 (2017). Article ADS CAS Google
Scholar * Sun, J.-K. et al. General synthetic route toward highly dispersed metal clusters enabled by poly(ionic liquid)s. _J. Am. Chem. Soc._ 139, 8971–8976 (2017). Article CAS Google
Scholar * Cheng, N. et al. Platinum single-atom and cluster catalysis of the hydrogen evolution reaction. _Nat. Commun._ 7, 13638 (2016). Article ADS CAS Google Scholar * Liu, J.
Catalysis by supported single metal atoms. _ACS Catal._ 7, 34–59 (2017). Article CAS Google Scholar * Li, Y. et al. A Pt-cluster-based heterogeneous catalyst for homogeneous catalytic
reactions: X-ray absorption spectroscopy and reaction kinetic studies of their activity and stability against leaching. _J. Am. Chem. Soc._ 133, 13527–13533 (2011). Article CAS Google
Scholar * Papp, C. Catalysis at the limit. _Nat. Chem._ 10, 995–996 (2018). Article CAS Google Scholar * Imaoka, T. et al. Platinum clusters with precise numbers of atoms for
preparative-scale catalysis. _Nat. Commun._ 8, 688 (2017). Article ADS Google Scholar * Tian, S. et al. Carbon nitride supported Fe2 cluster catalysts with superior performance for alkene
epoxidation. _Nat. Commun._ 9, 2353 (2018). Article ADS Google Scholar * Liu, L. & Corma, A. Metal catalysts for heterogeneous catalysis: from single atoms to nanoclusters and
nanoparticles. _Chem. Rev._ 118, 4981–5079 (2018). Article CAS Google Scholar * Therrien, A. J. et al. An atomic-scale view of single-site Pt catalysis for low-temperature CO oxidation.
_Nat. Catal._ 1, 192–198 (2018). Article CAS Google Scholar * Du, X. & Jin, R. Atomically precise metal nanoclusters for catalysis. _ACS Nano_ 13, 7383–7387 (2019). Article CAS
Google Scholar * Freund, H.-J. The surface science of catalysis and more, using ultrathin oxide films as templates: a perspective. _J. Am. Chem. Soc._ 138, 8985–8996 (2016). Article CAS
Google Scholar * Ishida, Y., Corpuz, R. D. & Yonezawa, T. Matrix sputtering method: a novel physical approach for photoluminescent noble metal nanoclusters. _Acc. Chem. Res._ 50,
2986–2995 (2017). Article CAS Google Scholar * Luza, L. et al. Revealing hydrogenation reaction pathways on naked gold nanoparticles. _ACS Catal._ 7, 2791–2799 (2017). Article CAS
Google Scholar * Torimoto, T. et al. Sputter deposition onto ionic liquids: simple and clean synthesis of highly dispersed ultrafine metal nanoparticles. _Appl. Phys. Lett._ 89, 243117
(2006). Article ADS Google Scholar * Deng, L. et al. Highly correlated size and composition of Pt/Au alloy nanoparticles via magnetron sputtering onto liquid. _Langmuir_ 36, 3004–3015
(2020). Article CAS Google Scholar * Pongthawornsakun, B. et al. Deposition of Pt nanoparticles on TiO2 by pulsed direct current magnetron sputtering for selective hydrogenation of
vanillin to vanillyl alcohol. _Catal. Today_ 358, 51–59 (2020). * Calabria, L. et al. Confined naked gold nanoparticles in ionic liquid films. _Nanoscale_ 9, 18753–18758 (2017). Article CAS
Google Scholar * Wender, H., Migowski, P., Feil, A. F., Teixeira, S. R. & Dupont, J. Sputtering deposition of nanoparticles onto liquid substrates: recent advances and future trends.
_Coord. Chem. Rev._ 257, 2468–2483 (2013). Article CAS Google Scholar * Eshetu, G. G., Armand, M., Ohno, H., Scrosati, B. & Passerini, S. Ionic liquids as tailored media for the
synthesis and processing of energy conversion materials. _Energy Environ. Sci._ 9, 49–61 (2016). Article CAS Google Scholar * König, D., Richter, K., Siegel, A., Mudring, A.-V. &
Ludwig, A. High-throughput fabrication of Au–Cu nanoparticle libraries by combinatorial sputtering in ionic liquids. _Adv. Funct. Mater._ 24, 2049–2056 (2014). Article Google Scholar *
Torimoto, T. et al. Ultrathin oxide shell coating of metal nanoparticles using ionic liquid/metal sputtering. _J. Mater. Chem. A_ 3, 6177–6186 (2015). Article CAS Google Scholar * Wender,
H. et al. Synthesis of gold nanoparticles in a biocompatible fluid from sputtering deposition onto castor oil. _Chem. Commun._ 46, 7019–7021 (2010). Article CAS Google Scholar * Sugioka,
D., Kameyama, T., Kuwabata, S., Yamamoto, T. & Torimoto, T. Formation of a Pt-decorated Au nanoparticle monolayer floating on an ionic liquid by the ionic liquid/metal sputtering method
and tunable electrocatalytic activities of the resulting monolayer. _ACS Appl. Mater. Interfaces_ 8, 10874–10883 (2016). Article CAS Google Scholar * Wender, H. et al. Sputtering onto
liquids: from thin films to nanoparticles. _J. Phys. Chem. C_ 115, 16362–16367 (2011). Article CAS Google Scholar * Kuwabata, S., Tsuda, T. & Torimoto, T. Room-temperature ionic
liquid. A new medium for material production and analyses under vacuum conditions. _J. Phys. Chem. Lett._ 1, 3177–3188 (2010). Article CAS Google Scholar * Cook, S. C., Padmos, J. D.
& Zhang, P. Surface structural characteristics and tunable electronic properties of wet-chemically prepared Pd nanoparticles. _J. Chem. Phys._ 128, 154705 (2008). Article ADS CAS
Google Scholar * Ohba, T. et al. An origin for lattice expansion in PVP-protected small Pd metal nanoparticles. _Bull. Chem. Soc. Jpn._ 90, 720–727 (2017). Article CAS Google Scholar *
Lakshmi Kantam, M. et al. Alumina supported copper nanoparticles for aziridination and cyclopropanation reactions. _Catal. Commun._ 8, 1963–1968 (2007). Article Google Scholar * Chen, S.,
Ma, J. & Wang, J. Palladium-catalyzed cyclopropanation of electron-deficient olefins with aryldiazocarbonyl compounds. _Tetrahedron Lett._ 49, 6781–6783 (2008). Article CAS Google
Scholar * Shishilov, O. N. et al. Palladium(II) carboxylates and palladium(I) carbonyl carboxylate complexes as catalysts for olefin cyclopropanation with ethyl diazoacetate. _Dalton
Trans_. 6626–6633, (2009). * Martín, C., Molina, F., Alvarez, E. & Belderrain, T. R. Stable N-heterocyclic carbene (NHC)–palladium(0) complexes as active catalysts for olefin
cyclopropanation reactions with ethyl diazoacetate. _Chem. Eur. J._ 17, 14885–14895 (2011). Article Google Scholar * Gärtner, D., Welther, A., Rad, B. R., Wolf, R. & Jacobi von
Wangelin, A. Heteroatom-free arene-cobalt and arene-iron catalysts for hydrogenations. _Angew. Chem. Int. Ed._ 53, 3722–3726 (2014). Article Google Scholar * Hamasaka, G., Ichii, S. &
Uozumi, Y. A palladium NNC-pincer complex as an efficient catalyst precursor for the Mizoroki−Heck reaction. _Adv. Synth. Catal._ 360, 1833–1840 (2018). Article CAS Google Scholar * Luo,
X. L. & Crabtree, R. H. Homogeneous catalysis of silane alcoholysis via nucleophilic attack by the alcohol on an Ir(_ƞ_2-HSiR3) intermediate catalyzed by [IrH2S2(PPh3)2]SbF6 (S =
solvent). _J. Am. Chem. Soc._ 111, 2527–2535 (1989). Article CAS Google Scholar * Neate, P. G. N., Greenhalgh, M. D., Brennessel, W. W., Thomas, S. P. & Neidig, M. L. Mechanism of the
bis(imino)pyridine-iron-catalyzed hydromagnesiation of styrene derivatives. _J. Am. Chem. Soc._ 141, 10099–10108 (2019). Article CAS Google Scholar * Burés, J. Variable time
normalization analysis: general graphical elucidation of reaction orders from concentration profiles. _Angew. Chem. Int. Ed._ 55, 16084–16087 (2016). Article Google Scholar * Burés, J. A
simple graphical method to determine the order in catalyst. _Angew. Chem. Int. Ed._ 55, 2028–2031 (2016). Article Google Scholar * Nielsen, C. D. T. & Burés, J. Visual kinetic
analysis. _Chem. Sci._ 10, 348–353 (2019). Article CAS Google Scholar * Neese, F. Software update: the ORCA program system, version 4.0. _WIREs Comput. Mol. Sci._ 8, e1327 (2018). Article
Google Scholar * Cho, D.-J. et al. Chiral C2-symmetric bisferrocenyldiamines as ligands for transition metal catalyzed asymmetric cyclopropanation and aziridination. _Tetrahedron_ 10,
3833–3848 (1999). Article CAS Google Scholar Download references ACKNOWLEDGEMENTS The authors thank the University of Nottingham Advanced Molecular Materials Interdisciplinary Research
Cluster, Propulsion Futures and Green Chemicals Beacons of Excellence for the financial support, the Nanoscale and Microscale Research Centre (nmRC) for access to materials characterisation
equipment and the National Mass Spectrometry Facility at Swansea University and Mark F. Wyatt. The authors acknowledge support of EPSRC CDT in Sustainable Chemistry (EP/L015633/1), EPSRC
projects (EP/K005138/1) and BBSRC (BB/L013940/1) for the financial support. E.H.Å. thanks the funding from Eemil Aaltonen Foundation and the access to the UoN’s Augusta HPC service. The
authors acknowledge the Diamond Light Source and the UK Catalysis Hub for provision of beam time (proposal number SP19850-5 and through the Block Allocation Group (BAG) for Energy Materials
under proposal sp17198) and Alan Chadwick and Giannantonio Cibin for assistance with the XAS measurements. UK Catalysis Hub provided resources and support funded by EPSRC grants:
EP/R026939/1, EP/R026815/1, EP/R026645/1, EP/R027129/1, and EP/M013219/1. E.J. thanks the Swedish Research Council for their financial support in the form of an International Postdoc
fellowship and Professor Clare Grey for the use of the Odyssey cluster for DFT calculations. The authors thank Mark Guyler (University of Nottingham) for the help with the magnetron
sputtering laboratory set-up and Dr Michael Fay and Dr Emily Smith (University of Nottingham) for the helpful scientific discussions. AUTHOR INFORMATION AUTHORS AND AFFILIATIONS * School of
Chemistry, University of Nottingham, Nottingham, UK Israel Cano, Andreas Weilhard, Carmen Martin, Jose Pinto, Rhys W. Lodge, Graham A. Rance, Elina Harriet Åhlgren, Andrei N. Khlobystov
& Jesum Alves Fernandes * GSK Carbon Neutral Laboratories for Sustainable Chemistry, University of Nottingham, Nottingham, UK Jose Pinto, Ana R. Santos & Peter Licence * Nanoscale
and Microscale Research Centre, University of Nottingham, Nottingham, UK Graham A. Rance * Department of Chemistry, University of Cambridge, Cambridge, UK Erlendur Jónsson * Department of
Physics, Chalmers University of Technology, Gothenburg, Sweden Erlendur Jónsson * Department of Physics, University of York, York, UK Jun Yuan * Nanoscale Physics Research Laboratory, School
of Physics and Astronomy, University of Birmingham, Birmingham, UK Ziyou Y. Li Authors * Israel Cano View author publications You can also search for this author inPubMed Google Scholar *
Andreas Weilhard View author publications You can also search for this author inPubMed Google Scholar * Carmen Martin View author publications You can also search for this author inPubMed
Google Scholar * Jose Pinto View author publications You can also search for this author inPubMed Google Scholar * Rhys W. Lodge View author publications You can also search for this author
inPubMed Google Scholar * Ana R. Santos View author publications You can also search for this author inPubMed Google Scholar * Graham A. Rance View author publications You can also search
for this author inPubMed Google Scholar * Elina Harriet Åhlgren View author publications You can also search for this author inPubMed Google Scholar * Erlendur Jónsson View author
publications You can also search for this author inPubMed Google Scholar * Jun Yuan View author publications You can also search for this author inPubMed Google Scholar * Ziyou Y. Li View
author publications You can also search for this author inPubMed Google Scholar * Peter Licence View author publications You can also search for this author inPubMed Google Scholar * Andrei
N. Khlobystov View author publications You can also search for this author inPubMed Google Scholar * Jesum Alves Fernandes View author publications You can also search for this author
inPubMed Google Scholar CONTRIBUTIONS I.C., A.W. and C.M. performed part of the ionic liquid synthesis and catalytic experiments. R.W.L., J.Y. and Z.Y.L. conducted TEM and AC-STEM
measurements. J.P. performed the magnetron sputtering deposition of Pd species in ionic liquid and ICP-OES measurements. A.S. performed part of the ionic liquid synthesis and XPS
measurements. G.A.R. conducted the Raman spectroscopy measurements. E.H.Å. conducted the classical molecular dynamics simulations with the PARCAS code. E.J. performed the DFT calculations.
P.L., A.N.K. and J.A.F. designed the study, analysed the data and co-wrote the paper. All the authors discussed the results and commented on the manuscript. CORRESPONDING AUTHOR
Correspondence to Jesum Alves Fernandes. ETHICS DECLARATIONS COMPETING INTERESTS The authors declare no competing interests. ADDITIONAL INFORMATION PEER REVIEW INFORMATION _Nature
Communications_ thanks the anonymous reviewers for their contribution to the peer review of this work. Peer reviewer reports are available. PUBLISHER’S NOTE Springer Nature remains neutral
with regard to jurisdictional claims in published maps and institutional affiliations. SUPPLEMENTARY INFORMATION SUPPLEMENTARY INFORMATION PEER REVIEW FILE RIGHTS AND PERMISSIONS OPEN ACCESS
This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as
long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The images or other third
party material in this article are included in the article’s Creative Commons license, unless indicated otherwise in a credit line to the material. If material is not included in the
article’s Creative Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright
holder. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/. Reprints and permissions ABOUT THIS ARTICLE CITE THIS ARTICLE Cano, I., Weilhard, A., Martin, C.
_et al._ Blurring the boundary between homogenous and heterogeneous catalysis using palladium nanoclusters with dynamic surfaces. _Nat Commun_ 12, 4965 (2021).
https://doi.org/10.1038/s41467-021-25263-6 Download citation * Received: 19 March 2020 * Accepted: 29 July 2021 * Published: 17 August 2021 * DOI: https://doi.org/10.1038/s41467-021-25263-6
SHARE THIS ARTICLE Anyone you share the following link with will be able to read this content: Get shareable link Sorry, a shareable link is not currently available for this article. Copy to
clipboard Provided by the Springer Nature SharedIt content-sharing initiative