Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
View Article Online View Journal Catalysis Science & Technology Accepted Manuscript This article can be cited before page numbers have been issued, to do this please use: N. Yuan, M. Majeed, É. G. Bajnóczi , A. R. Persson, L. R. Wallenberg, A. K. Inge, N. Heidenreich, N. Stock, X. Zou, O. F. Wendt and I. Persson, Catal. Sci. Technol., 2019, DOI: 10.1039/C8CY02430H. Volume 6 Number 1 7 January 2016 Pages 1–308 Catalysis Science & Technology www.rsc.org/catalysis This is an Accepted Manuscript, which has been through the Royal Society of Chemistry peer review process and has been accepted for publication. Accepted Manuscripts are published online shortly after acceptance, before technical editing, formatting and proof reading. Using this free service, authors can make their results available to the community, in citable form, before we publish the edited article. We will replace this Accepted Manuscript with the edited and formatted Advance Article as soon as it is available. You can find more information about Accepted Manuscripts in the author guidelines. ISSN 2044-4753 PAPER Qingzhu Zhang et al. Catalytic mechanism of C–F bond cleavage: insights from QM/MM analysis of fluoroacetate dehalogenase Please note that technical editing may introduce minor changes to the text and/or graphics, which may alter content. The journal’s standard Terms & Conditions and the ethical guidelines, outlined in our author and reviewer resource centre, still apply. In no event shall the Royal Society of Chemistry be held responsible for any errors or omissions in this Accepted Manuscript or any consequences arising from the use of any information it contains. rsc.li/catalysis Please doScience not adjust margins Catalysis & Technology Page 1 of 8 View Article Online Journal Name Open Access Article. Published on 25 March 2019. Downloaded on 3/29/2019 10:27:03 AM. This article is licensed under a Creative Commons Attribution 3.0 Unported Licence. ARTICLE Received 00th January 20xx, Accepted 00th January 20xx DOI: 10.1039/x0xx00000x www.rsc.org/ In situ XAS study of the local structure and oxidation state evolutions of palladium in a reduced graphene oxide supported Pd(II) carbene complex during an undirected C−H acetoxylation reaction† Ning Yuan,*†a,b Maitham H. Majeed,†c Éva G. Bajnóczi,a Axel R. Persson,c,d L. Reine Wallenberg,c,d A. Ken Inge,b Niclas Heidenreich,e,f Norbert Stock,e Xiaodong Zou,b Ola F. Wendt*c and Ingmar Persson*a In situ X-ray absorption spectroscopy (XAS) investigations have been performed to provide insights into the reaction mechanism of a palladium(II) catalyzed undirected C–H acetoxylation reaction in the presence of an oxidant. A Pd(II) Nheterocyclic carbene complex -stacked onto reduced graphene oxide (rGO) was used as catalyst. The Pd speciation during the catalytic process was examined by XAS, which revealed a possible mechanism over the course of the reaction. Pd(II) complexes in the as-synthesized catalyst first go through a gradual ligand substitution where chloride ions bound to Pd(II) are replaced by other ligands with a bond distance to Pd corresponding to carbon, nitrogen and/or oxygen (L). Parallel to this the mean oxidation state of Pd increases indicating the formation of Pd(IV) species. At a later stage, a fraction of the Pd complexes start to slowly transform into Pd nanoclusters. The mean average oxidation state of Pd decreases to the initial state at the end of the experiment which means that comparable amounts of Pd(0) and Pd(IV) are present. These observations from heterogeneous catalysis are in good agreement with its homogeneous analog and they support a Pd(II)Pd(IV)-Pd(II) reaction mechanism. Introduction The formation of carbon–oxygen (C–O) bonds directly from carbon–hydrogen (C–H) bonds is an attractive synthetic pathway to many important organic compounds such as pharmaceuticals and agrochemicals.1,2 Such transformations are usually catalyzed by transition metals (TMs) and most examples include homogeneous catalysts, but heterogeneous TM catalysts are in demand, not only because these supported systems are recyclable and easy to separate from the reaction mixture, but also for their potential to introduce novel chemical reactivity.3,4,5 Recently, reduced graphene oxide (rGO) was successfully applied as a supporting material for a. Department of Molecular Sciences, Swedish University of Agricultural Sciences, P.O. Box 7015, SE-750 07 Uppsala, Sweden. E-mail: ning.yuan@mmk.su.se of Materials and Environmental Chemistry, Stockholm University, SE106 91 Stockholm, Sweden. E-mail: ingmar.persson@slu.se c. Centre for Analysis and Synthesis, Department of Chemistry, Lund University, P.O. Box 124, SE-221 00 Lund, Sweden. E-mail: ola.wendt@chem.lu.se d. National Center for High Resolution Electron Microscopy and NanoLund, Lund University, Box 124, SE-221 00 Lund, Sweden e. Institut für Anorganische Chemie, Christian-Albrechts-Universität zu Kiel, DE24118 Kiel, Germany f. Deutsches-Elektronen-Synchrotron DESY, DE-22607 Hamburg, Germany † These authors contributed equally to this work. † Electronic Supplementary Information (ESI) available: Details about XAS data collection and analysis, XANES spectra of the complex under study, free and bound to rGO, and EXAFS data treatment of data collected in in situ mode See DOI: 10.1039/x0xx00000x b. Department different types of transition metal N-heterocyclic carbene (TM–NHC) complexes and applied for many transformations including C−H acetoxylation.6–10 Recently, we developed an anthracene-tagged Pd(II)–NHC complex, 1, supported on rGO through -stacking (here labeled 1@rGO) according to an adaptation of the procedure proposed by Peris and coworkers.8 The preparation procedure is depicted in Scheme 1.11 It was shown that 1@rGO works as an active heterogeneous catalyst in undirected C−H acetoxylation of benzene and its catalytic efficiency is slightly higher than its homogenous analog.11 This, together with our previous work,4 shows that Pd(II)–NHC complexes are effective as supported catalysts for arene acetoxylation. Ex situ studies indicate that the catalyst is not reduced during catalysis but there is little information on the mechanism.4,11 It can be noted that there are also systems based on Pd(0) pre-catalysts but these were applied for directed C–H activation.3 Although a previous study by Tato et al. has shed light on our understanding of the reaction mechanism of C−H acetoxylation catalyzed by Pd(II)–NHC complexes,12 such efforts were limited to homogenous reaction conditions due to the complexity of heterogeneous systems. Also, the oxidation state evolution during directed catalysis with Pd nanoparticles has been studied.13 However, there are few, if any, experimental studies on the nature and particularly oxidation state of the Pd catalytic species during undirected C–H This journal is © The Royal Society of Chemistry 20xx J. Name., 2013, 00, 1-3 | 1 Please do not adjust margins Catalysis Science & Technology Accepted Manuscript DOI: 10.1039/C8CY02430H Please doScience not adjust margins Catalysis & Technology Page 2 of 8 Journal Name activation. Many of the traditional techniques used for characterization of catalysts are either difficult to conduct under reaction conditions, such as X-ray photoelectron spectroscopy (XPS), or unable to probe the catalytic centres directly, such as Fourier-transform infrared (FTIR) spectroscopy. To meet the demand of studying the catalytic species in a direct manner on atomic level, X-ray absorption spectroscopy (XAS) has become a suitable experimental approach, as it is element specific with high sensitivity to local structure and valence state of the absorbing element.14 XAS can be applied on all states of aggregation as well as their mixtures, and only millimolar concentrations are required for successful structure analysis around the absorbing atom. Furthermore, a scanning rate of minutes allows in situ XAS measurements to be conducted under various chemical conditions.15–17 In the Scheme 1. Immobilization of 1 on rGO. recent years of in situ XAS research on catalysis, major efforts have been focussed on solid-gas heterogeneous catalytic systems due to the manageable design and construction of in situ reactors.18–23 For solid-liquid heterogeneous reactions or liquid homogenous reactions, it becomes more challenging to build up a suitable sample cell because of more demanding reaction conditions and interference from the reaction mixture. The growing interest in in situ XAS measurements on solid-liquid and solution phase reactions has promoted the development of the reactors.24,25 With improved control of the chemical conditions several successful examples have been reported.26–29 These studies have proven that in situ XAS is a powerful tool to study the catalytic species in solution and solution-solid reaction mixtures. In the current study we aim to elucidate the palladium species and their oxidation states over the course of the undirected C−H acetoxylation of benzene catalyzed by 1@rGO using in situ XAS spectroscopy and gain insights into the reaction mechanism. The focus is on the heterogeneous system and the homogeneous analog is used for comparison. Experimental section Materials. 1@rGO (2.5 wt% Pd loading) and 1 were used as catalysts and synthesized according to procedures reported previously.8,11 All reagents and solvents used in the experiments were purchased from commercial suppliers without further purification. Catalytic reactions for in situ XAS experiments. Benzene (1.75 g, 22.44 mmol, 21.77 equiv.), (Diacetoxyiodo)benzene (oxidant, 0.33 g, 1.03 mmol, 1 equiv.), 1@rGO (0.10 g, 2.3 mol% with respect to the oxidant) and 1 (0.07 g, 9.2 with Viewmol% Article Online DOI: 10.1039/C8CY02430H respect to the oxidant), glacial acetic acid (1.50 mL), and acetic anhydride (0.17 mL) were transferred into a reaction vessel which was sealed in the preparation room and transported immediately to the hutch for measurements. The reaction mixture was stirred at 92 °C during the in situ XAS measurements. Reactor for in situ experiments. In situ XAS data were collected using a custom-made reactor designed for investigations under solvothermal conditions.24 The remotely controlled reactor was used to heat and stir the reaction mixture in a sealed 5mL glass vial with 1 mm thick walls, allowing the synchrotron radiation to penetrate the sample in transmission mode. The target temperature is typically achieved within two minutes, and a thermocouple was immersed into the reaction mixture to monitor the temperature throughout the experiment. Figure S1 in Section S1 shows the exploded-view illustration of the reactor. XAS experiments. All XAS data were collected at beamline P64 at Petra III Extension, Deutsches Elektronen-Synchrotron (DESY), Hamburg, Germany. The XAS measurements were conducted in transmission mode at the Pd K-edge (24.349 keV) with an energy range from 24.00 to 25.15 keV for dry samples or 25.00 keV for in situ measurement.30 A palladium metal foil was measured simultaneously and its first inflection point on the absorption edge was used to individually calibrate all the XAS spectra during the measurements. As-synthesized 1@rGO, 1, and recycled 1@rGO were measured using a standard sample holder for powders. For the in situ measurements, each XAS scan was set to ca. 6 min with considerations of both time resolution and data quality. More technical details and data analysis are described in Section S1. Results and discussion The successful preparation of 1@rGO is confirmed by both X-ray absorption near edge structure (XANES) and extended Xray absorption fine structure (EXAFS) spectra of the assynthesized 1 and 1@rGO (Figure S2-a in Section S2 and Figure S3 in Section S3). The structure of 1 was described in the previous work where each Pd binds two chloride ions (Cl−), one carbene and one pyridine ligand. The oxidation state of Pd was determined to +II.11 This information is consistent with the analysis of the XANES spectra of 1 and its structure determined by EXAFS (Table 1). The XANES and EXAFS spectra of 1@rGO are similar to those of 1 indicating a successful immobilization of 1 onto rGO. EXAFS refinement of 1@rGO reveals one Pd–C bond with a length of 1.95 Å which corresponds to the bond between Pd and the NHC ligand. Another bond distance to Pd with a relatively large error, 0.08 Å, is observed at 2.18 Å. This bond length is close to the distance between Pd and pyridine ligand in 1. However, our previous results based on XPS show that the pyridine ligand dissociates from Pd and is stacked on the support.11 One possibility is that Pd atoms are pi-bound to a C=C double bond and 2.18 Å is a reasonable bond length in 2 | J. Name., 2012, 00, 1-3 This journal is © The Royal Society of Chemistry 20xx Please do not adjust margins Catalysis Science & Technology Accepted Manuscript Open Access Article. Published on 25 March 2019. Downloaded on 3/29/2019 10:27:03 AM. This article is licensed under a Creative Commons Attribution 3.0 Unported Licence. ARTICLE Catalysis & Technology Please doScience not adjust margins Journal Name ARTICLE Open Access Article. Published on 25 March 2019. Downloaded on 3/29/2019 10:27:03 AM. This article is licensed under a Creative Commons Attribution 3.0 Unported Licence. such configuration.31–33 Scheme 1 describes the immobilization process based on our understanding. In the previous study, we have shown that 1@rGO is an effective catalyst in the acetoxylation of benzene and a few other arenes. The reaction conditions were optimized and used for the in situ XAS measurement as described in Scheme 2. Benzene was chosen as a model substrate with excess amount of oxidant PhI(OAc)2. form of Pd(III) species is dimers.37,38 Interestingly, the first in View Article Online 10.1039/C8CY02430H situ XANES spectrum displayed changesDOI: in comparison to assynthesized 1@rGO, and thereafter the observed changes slow down significantly. This implies a process of pre-catalyst activation in the beginning of the reaction which has been realized as an important procedure for the catalyst which will be further discussed in the EXAFS analysis.29,39 Scheme 2. 1@rGO catalyzed undirected C–H acetoxylation reaction of benzene for in situ measurement. 1@rGO was added to the reaction mixture and the XAS data collection started when the reaction was initiated with a time resolution of ca. 6 min. Due to the low concentration of Pd (ca. 7 mM) and low signal/noise (S/N) ratio, groups of XAS spectra with identical features were averaged to improve the statistics. Figure 1 displays the selected Pd K-edge XANES spectra during the in situ measurement which is divided into two parts. The first part of the measurement is from the beginning of the reaction to ca. 80 min as presented in Figure 1a. The main changes of the XANES spectra captured here are the increased white line intensity, especially during the first six minutes of the reaction, and a continuous shift of edge position towards higher energies. The increase in white line intensity has been proposed to be caused by an increase in the oxidation state of palladium in molecular complexes.34,35 Regarding the edge shift observation, theoretically, it can be caused either by a partial oxidation of Pd(II) to Pd(IV), or a change of bound ligands from covalently bound to more electrostatically bound ones. In this system the only strong electron-pair donor to Pd in 1@rGO is the NHC ligand and the Pd–NHC bond is very stable with a calculated bond energy of 45-50 kcal·mol-1.36 Therefore, it is expected to remain bound to Pd during the reaction. Furthermore, the 1@rGO catalyst showed good recyclability and could be reused up to four runs with only slight decrease in the activity which attests to the stable Pd–C(NHC) bond. Hence, a partial oxidation of Pd(II) complexes seems to take place by the shift towards higher energy, together with the increased white line intensity. Even though the edge shift is small, ca. 1 eV, and corresponds to a mean oxidation state change less than one, it is clearly observable. However, XAS spectra display a mean value of the radiated absorbing species, and a shift of ca. 1 eV, ca. 25% of an expected shift between Pd(II) and Pd(IV), implies a reasonable scenario where a minor fraction of the Pd(II) complexes have been oxidized to Pd(IV), considering the relatively short lifetime of Pd(IV) in the catalytic redox cycle. The argument to propose Pd(IV) instead of Pd(III) is that no back Fourier transformed EXAFS spectra indicating Pd---Pd dimers with short Pd---Pd bonds were observed; the common Figure 1. In situ Pd K-edge XANES spectra of 1@rGO catalyzed C–H acetoxylation reaction. (a) First part of measurement when edge position shifts towards higher energy. (b) Second part of measurement when edge position shifts towards lower energy. The edge shifts are magnified as inset. Selected XANES spectra of the second part of the in situ measurement, 80 min - 24 h, are given in Figure 1b. As the measurement proceeds, the edge position shifts backward to lower energies. The shift continues until the end of the measurement at 24 h. The XANES spectrum of the recycled 1@rGO has an edge position and shape after the edge identical to the catalyst at 24 h. This indicates that the evolution of the Pd species ceases after 24 h, and that the recycling procedure does not introduce further changes of the Pd species. It is noted that the edge position of Pd at the end of the measurement returns to a position very close to the one of as-synthesized 1@rGO (Figure 1 and Figure S2-b in Section S2). This means that the mean oxidation state of Pd at the end of the measurement is +II. Meanwhile, the white line intensity has a general trend of declining even though the S/N ratio is This journal is © The Royal Society of Chemistry 20xx J. Name., 2013, 00, 1-3 | 3 Please do not adjust margins Catalysis Science & Technology Accepted Manuscript Page 3 of 8 Please doScience not adjust margins Catalysis & Technology Journal Name Open Access Article. Published on 25 March 2019. Downloaded on 3/29/2019 10:27:03 AM. This article is licensed under a Creative Commons Attribution 3.0 Unported Licence. low. The shape of the absorption edge appears to more resemble that of metallic Pd even though it is damped due to the small fraction. Figure 2. Fourier transformed k3-weighted EXAFS data as a function of measurement time showing the coordination environment of Pd in 1@rGO and 1. The spectra are not phase correct and the k range used is fixed at 2-9 Å-1 for comparison. The arrows indicate the stepwise dissociation of Cl- ligands. These observations indicate that the mean oxidation state of the Pd species changes in the opposite direction in the second part as compared to the first one, and that the Pd(IV) species is mostly reduced to Pd(II) and to a small extent to Pd(0) as the reaction proceeds to the end. Interestingly, a small fraction of Pd(IV) species should remain considering the mean oxidation state of +II. Kim and co-workers have shown the co-existence of Pd(0), Pd(II) and Pd(IV) species by oxidizing Pd nanoparticles at controlled conditions where Pd(IV) formed polymeric species with enhanced stability.13 Although such Pd(IV) signals are not observed in the EXAFS of our material, it can be attributed to its low content. The specific coordination environment of the Pd species in 1@rGO during the reaction were also studied by EXAFS and the Fourier transformed spectra are presented in Figure 2a. The k range of all the EXAFS spectra is 2-9 Å−1 for easier comparison of the data at different reaction stages. The EXAFS spectra in full treatable range and their refinements are reported in Table 1 and Figure S4 in Section S3. From the assynthesized 1@rGO to 6 min an obvious shift of the main peak towards shorter distance is observed indicating a change in the coordination environment of Pd, Figure 2a. The refinement of the EXAFS spectrum of the catalyst at 6 min shows that on average each Pd centre is bound to three L ligands (L ligands: C, N and/or O ligands) at a mean distance of ca. 2.01 Å and one Cl− ligand at ca. 2.32 Å. Because of the average nature of EXAFS the refinement parameters can be also interpreted as half of the palladium centres have four coordinated L ligands and no Cl− ligands, while the other half retain the coordination structure of the as-synthesized catalyst. However, the second View Article Online 10.1039/C8CY02430H scenario does not seem reasonable DOI: considering the local structural change of Pd slows down immediately after 6 min and certain transformations should be completed. As the changes of the XANES spectra after 6 min in Figure 1a are less pronounced, the differences of EXAFS spectra from 6 to 80 min are also very small and cannot be resolved by refinements at the current resolution. This implies that the local structure of the palladium centre remains the same after activation while a fraction of them have been further oxidized and bound to ligands with a Pd–L bond length of ca. 2.0 Å. Another change occurs from 80 min to 3.5 h where the main peak moves further towards shorter distances. The refinement reveals that the remaining Cl− is replaced by another L ligand resulting in Pd complexes with four L ligands after 3.5 h. At the same time, the introduction of a Pd–Pd bond distance of 2.7 Å corresponding to Pd nanoclusters with a mean coordination number of ca. 0.8 is necessary to achieve a proper fit of the EXAFS spectrum. This signal can be observed as the relatively small peak at ca. 2.3 Å (without phase correction) after the main peak. It strongly implies that a small fraction of the Pd complexes have been reduced to metallic Pd nanoclusters; this could be caused by a reductive elimination of ligands. Interestingly, the contribution of Pd–Pd bond distances is not rapidly growing between 6 and 8 hours, and the local environment of Pd seems to be stable. This observation is in contrast to a previously investigated system where the transformation of mononuclear Pd complexes to Pd nanoclusters is much faster.29 Although the local structure of the Pd appears to be the same during this process, the nature of the L ligands bound to Pd seems to change as the edge position shifts towards lower energies. At 24 h the peak at 2.7 Å (after phase correction) became more pronounced indicating an increased amount of metallic Pd nanoclusters. Meanwhile, the first main peak shifts towards longer distance at 24 h. It is known that the Pd–Pd bond has a satellite peak with similar appearance as a Pd–Cl bond and it becomes more pronounced when the applied k range is shortened.29 Due to limited EXAFS data quality for the catalyst at 24 h, the satellite peak of Pd–Pd bond distances and the Pd–L peak merge causing the abnormal peak position. The spectrum of the recycled catalyst is similar to the one at 24 h and ca. 40% of Pd is estimated to be present as Pd nanoclusters based on linear combination fit of the XANES spectrum (Figure S6a, Section S4), while the remaining palladium is in the form of Pd(II) complexes with a small amount of Pd(IV) complexes. The Pd species in the recycled 1@rGO are further proven by transmission electron microscopy (TEM) and X-ray energy dispersive spectroscopy (EDS) showing the presence of Pd well dispersed over the whole observing area of the catalyst, including Pd aggregates (Section S5). The oxidant is thermally degraded over time which means that towards the end of the reaction there is no or little oxidant left.4 In the absence of oxidant Pd may be reduced which could lead to the formation of metallic Pd nanoclusters although these were not observed in the previous ex situ investigation.11 4 | J. Name., 2012, 00, 1-3 This journal is © The Royal Society of Chemistry 20xx Please do not adjust margins Catalysis Science & Technology Accepted Manuscript ARTICLE Page 4 of 8 Catalysis & Technology Please doScience not adjust margins Page 5 of 8 ARTICLE Previously, it was also found that 1@rGO mostly retained its activity for at least four runs in the un-directed acetoxylation of benzene. However, the GC yield of acetoxybenzene showed a slow decreasing trend from 50% (first cycle) to 46% (fourth cycle). The solutions separated from the catalyst after each cycle contained less than 0.011 ppm.11 Since there is little Pd leaching, we propose that the slight decrease of productivity using recovered catalyst could be be related to the formation of Pd nanoclusters which should be an inactive species under the current reaction conditions. Although the oxidant is in excess in the in situ measurement, its relative amount is lower than in the reported catalysis conditions, meaning that the formation of Pd nanoclusters in this work is expected to be more extensive. The homogenous analog of 1@rGO was also investigated to compare with the observations and conclusions from 1@rGO. 1 was used as catalyst and the in situ experiment was repeated under the same condition as for 1@rGO except that the concentration of Pd in the homogenous reaction mixture was increased to ca. 35 mM to improve the absorbance and the S/N ratio of the individual in situ XAS scans. Figure 3 shows selected Pd K-edge XANES spectra of 1 during the reaction. The same strategy as for 1@rGO was used where the measurement is divided into two parts. The first part is from the beginning of the reaction to 78 min. It is noted that the white line intensity increases continuously until 52 min, while the edge position is shifting towards higher energy, a shift that also ceases at 52 min. There is no additional change from 52 to 78 min. From 78 min to 33 h after the start of the reaction, the white line intensity and the edge position return to the starting point. These observations are in good agreement with the observations using 1@rGO. The oxidation state of Pd in 1 also increases at first, followed by a decrease at the later stage of the measurement. View Article Online DOI: 10.1039/C8CY02430H Figure 3. In situ Pd K-edge XANES spectra of 1 catalyzed C–H acetoxylation reaction. (a) First part of measurement when edge position shifts towards higher energy. (b) Second part of measurement when edge position shifts towards lower energy. The edge shifts are magnified as insets. Table 1. Refined distances (d/Å), and mean number of distances (N) and Debye-Waller factor (σ2/Å2) in selected scans using 1@rGO and 1 as catalysts. Single scattering at outer shells and multiple scatterings are not shown in the table. Catalyst As-synth. 1@rGO 6 min (22, 80min) 3.5 h 8 h (6 h) Recycled (24 h) d(Pd-N/C/O) 1.953(6) 2.18(8) 2.013(6) 2.020(4) 2.015(7) 1.966(4) Nb 1.0 1.0 3.0 4.0 3.5 2.0 σ2 0.0017(7) 0.007(8) 0.009(1) 0.006(1) 0.005(2) 0.004(2) d(Pd-Cl) 2.293(9) N 2.0 σ2 0.0045(9) 2.320(6) 1.0 0.007(2) d(Pd-Pd) 2.68(2) 2.70(4) 2.652(4) N σ2 0.8 1.0 3.5 0.007(2) 0.005(4) 0.010(4) As-synth. 1 1.95(1) 1.0 0.003(2) 2.301(6) 2.0 0.0015(9) 2.17(8) 1.0 0.005(9) 6 min 2.00(1) 2.5 0.009(1) 2.293(8) 1.5 0.0023(7) 40 min 2.017(7) 3.0 0.003(1) 2.29(3) 1.0 0.003(1) 78 min 2.034(3) 4.0 0.0036(2) 2.65(2) 0.5 0.005(1) 5.5 h 2.032(2) 3.5 0.0043(4) 2.709(7) 1.0 0.010(1) 33 h 2.025(9) 3.0 0.007(2) 2.69(1) 3.0 0.012(2) aThe standard deviations in parentheses were obtained from k3 weighted least-squares refinements of the EXAFS function χ(k) and do not include systematic errors of the measurement. b The estimated error of the N values is ca. 25%. Underscored parameters have been optimized and fixed in the refinement. This journal is © The Royal Society of Chemistry 20xx J. Name., 2013, 00, 1-3 | 5 Please do not adjust margins Catalysis Science & Technology Accepted Manuscript Open Access Article. Published on 25 March 2019. Downloaded on 3/29/2019 10:27:03 AM. This article is licensed under a Creative Commons Attribution 3.0 Unported Licence. Journal Name Please doScience not adjust margins Catalysis & Technology Page 6 of 8 View Article Online Journal Name Open Access Article. Published on 25 March 2019. Downloaded on 3/29/2019 10:27:03 AM. This article is licensed under a Creative Commons Attribution 3.0 Unported Licence. ARTICLE The corresponding EXAFS spectra of 1 catalyzed reaction were analyzed as well, and their Fourier transformed spectra are displayed in Figure 2b. The k range used for 1 and 1@rGO is the same for proper comparison. Due to the better S/N ratio each individual in situ scan could be analyzed and they are presented as function of reaction time. In general, the change in local structure around Pd in 1 is the same as in 1@rGO. The first main peak gradually shifts towards shorter bond lengths corresponding to a stepwise dissociation of Cl− ligands, after which the peak corresponding to Pd–Pd single scattering appears and grows at a later stage of the measurement indicating the formation of metallic Pd nanoclusters. It is also noted that the intensity of the white line in Figure 3a increases more smoothly than in the case of 1@rGO. This might be a consequence of the higher Pd concentration in the 1 catalyzed reaction. However, the earlier catalysis study showed that 1@rGO lead to higher yield of acetoxybenzene compared to 1 and this observation might also be correlated to the slower activation rate of 1 in the beginning of the reaction.11 The results of the refinements of the EXAFS spectra are summarized in Table 1, and the fits are shown in Figure S5 in Section S3. The fraction of the metallic Pd nanoclusters at the end of the measurement (33 h) is estimated to be ca. 25% from the linear combination fit of its XANES spectrum (Figure S6b, Section S4). As outlined above this fraction is probably higher in the in situ measurements than in the actual catalysis because of the higher mol% of the catalyst with respect to the oxidant loading in the former. The fraction of Pd nanoclusters in 1 at 33 h is noted to be lower than in the recycled 1@rGO, which is also reflected by its lower mean number of Pd–Pd bonds shown in Table 1. remains, Pd(0) nanoclusters are slowly formed resulting in a relatively stable mixture of mononuclear Pd complexes and Pd metal nano-sized aggregates. During this process the mean oxidation state of Pd returns to the starting state resulting in a mixture dominated by the Pd(II) species, and smaller amounts of Pd(0) nanoclusters and Pd(IV) species. Overall, our observations provide evidence for a mechanism including oxidation state changes Pd(II)-Pd(IV)-Pd(II) during the undirected C−H acetoxylation reaction as proposed through computational methods.40,41 Conflicts of interest There are no conflicts to declare. Acknowledgements This work has been supported by the MATsynCELL project through Röntgen-Ångström Cluster, the Swedish Research Council (VR) and the German Federal Ministry of Education and Research (BMBF). We are also thankful to the Berzelii Center EXSELENT on Porous Materials, the Swedish Research Council (VR) and the Project Management Organization at DESY (Deutsches Elektronen-Synchrotron). The allocation of beamtime at P64, Petra III Extension, DESY is gratefully acknowledged. We thank the staff at beamline P64, particularly Dr. Vadim Murzin and Dr. Wolfgang Caliebe for the assistance during the data collection. To further improve the understanding of the activation of the pre-catalyst, the XAS spectrum of 1 in the reaction mixture at room temperature was collected and it is identical to the assynthesized 1. This demonstrates that the induced heat triggers the exchange of the ligands bound to Pd leading to the activation of the catalyst. 2 3 Conclusions 4 In conclusion, we have used in situ XAS to probe the speciation and the mean oxidation state of palladium over the course of an undirected C−H acetoxylation reaction. The heterogeneous catalyst 1@rGO and its homogenous analog 1 were examined. In both cases, a stepwise exchange of Cl− ions by L ligands (L: carbon, nitrogen and/or oxygen) is observed. Meanwhile a small fraction of Pd(II) is oxidized to Pd(IV). At a later stage of the reaction, when less or no oxidant (PhI(OAc)2) Notes and references 1 5 6 7 G. Dyker, Handbook of C-H transformations, Wiley-VCH, Weinheim, 2005. J.-Q. Yu and Z. Shi, C-H Activation, Springer Berlin, 2010. V. Pascanu, F. Carson, M. V. Solano, J. Su, X. Zou, M. J. Johansson and B. Martín-Matute, Chem. - A Eur. J., 2016, 22, 3729–3737. M. H. Majeed, P. Shayesteh, L. R. Wallenberg, A. R. Persson, N. Johansson, L. Ye, J. Schnadt and O. F. Wendt, Chem. - A Eur. J., 2017, 23, 8457–8465. E. Pérez-Mayoral, V. Calvino-Casilda and E. Soriano, Catal. Sci. Technol., 2016, 6, 1265–1291. D. Ventura-Espinosa, A. Carretero-Cerdán, M. Baya, H. García and J. A. Mata, Chem. - A Eur. J., 2017, 23, 10815– 10821. S. Sabater, J. A. Mata and E. Peris, Organometallics, 2015, 34, 1186–1190. This journal is © The Royal Society of Chemistry 20xx J. Name., 2013, 00, 1-3 | 6 Please do not adjust margins Catalysis Science & Technology Accepted Manuscript DOI: 10.1039/C8CY02430H Catalysis & Technology Please doScience not adjust margins Journal Name 8 9 10 11 12 Open Access Article. Published on 25 March 2019. Downloaded on 3/29/2019 10:27:03 AM. This article is licensed under a Creative Commons Attribution 3.0 Unported Licence. 13 14 15 16 17 18 19 20 21 22 23 24 25 ARTICLE S. Sabater, J. A. Mata and E. Peris, ACS Catal., 2014, 4, 2038–2047. S. Ruiz-Botella and E. Peris, Organometallics, 2014, 33, 5509–5516. S. Gonell and E. Peris, ACS Catal., 2014, 4, 2811–2817. M. H. Majeed, P. Shayesteh, A. R. Persson, L. R. Wallenberg, J. Schnadt and O. F. Wendt, Eur. J. Inorg. Chem., 2018, 4742–4746. F. Tato, A. Garcı́a-Domı́nguez and D. J. Cárdenas, Organometallics, 2013, 32, 7487–7494. K. Kim, Y. Jung, S. Lee, M. Kim, D. Shin, H. Byun, S. J. Cho, H. Song and H. Kim, Angew. Chemie Int. Ed., 2017, 56, 6952– 6956. L. Liu and A. Corma, Chem. Rev., 2018, 118, 4981–5079. J. Singh, C. Lamberti and J. A. Van Bokhoven, Chem. Soc. Rev., 2010, 39, 4754–4766. S. Bordiga, E. Groppo, G. Agostini, J. A. van Bokhoven and C. Lamberti, Chem. Rev., 2013, 113, 1736–1850. S. Bordiga, F. Bonino, K. P. Lillerud and C. Lamberti, Chem. Soc. Rev., 2010, 39, 4885–4927. S. K. Matam, M. H. Aguirre, A. Weidenkaff and D. Ferri, J. Phys. Chem. C, 2010, 114, 9439–9443. W. van Beek, O. V. Safonova, G. Wiker and H. Emerich, Phase Transitions, 2011, 84, 726–732. J.-D. Grunwaldt and B. S. Clausen, Top. Catal., 2002, 18, 37–43. S. Øien, G. Agostini, S. Svelle, E. Borfecchia, K. A. Lomachenko, L. Mino, E. Gallo, S. Bordiga, U. Olsbye, K. P. Lillerud and C. Lamberti, Chem. Mater., 2015, 27, 1042– 1056. J. D. Grunwaldt, M. Caravati, S. Hannemann and A. Baiker, Phys. Chem. Chem. Phys., 2004, 6, 3037–3047. C. Barzan, A. Piovano, L. Braglia, G. A. Martino, C. Lamberti, S. Bordiga and E. Groppo, J. Am. Chem. Soc., 2017, 139, 17064–17073. N. Heidenreich, U. Rütt, M. Köppen, A. K. Inge, S. Beier, A.C. Dippel, R. Suren and N. Stock, Rev. Sci. Instrum., 2017, 88, 104102. J.-D. Grunwaldt, M. Ramin, M. Rohr, A. Michailovski, G. R. Patzke and A. Baiker, Rev. Sci. Instrum., 2005, 76, 054104. 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 S. G. Fiddy, J. Evans, T. Neisius, M. A. Newton,View N. Tsoureas, Article Online A. A. D. Tulloch and A. A. Danopoulos, - A Eur. J., DOI:Chem. 10.1039/C8CY02430H 2007, 13, 3652–3659. P. J. Ellis, I. J. S. Fairlamb, S. F. J. Hackett, K. Wilson and A. F. Lee, Angew. Chemie Int. Ed., 2010, 49, 1820–1824. S. Reimann, J. Stötzel, R. Frahm, W. Kleist, J.-D. Grunwaldt and A. Baiker, J. Am. Chem. Soc., 2011, 133, 3921–3930. N. Yuan, V. Pascanu, Z. Huang, A. Valiente, N. Heidenreich, S. Leubner, A. K. Inge, J. Gaar, N. Stock, I. Persson, B. Martín-Matute and X. Zou, J. Am. Chem. Soc., 2018, 140, 8206–8217. M. . K. Thompson, A.; Attwood, D.; Gullikson, E.; Howells, P. . R. K.-J.; Kirz, J.; Kortright, J.; Lindau, I.; Liu, Y.; Pianetta and H. A.; Scofield, J.; Underwood, J.; Williams, G.; Winick, X-ray data booklet, Lawrence Berkeley National Laboratory: Berkeley, 3rd edn., 2009. K. T. Horak, S. Lin, J. Rittle and T. Agapie, Organometallics, 2015, 34, 4429–4432. S. Suseno and T. Agapie, Organometallics, 2013, 32, 3161– 3164. T. Murahashi, M. Fujimoto, M.-A. Oka, Y. Hashimoto, T. Uemura, Y. Tatsumi, Y. Nakao, A. Ikeda, S. Sakaki and H. Kurosawa, Science, 2006, 313, 1104–7. S. H. Eitel, M. Bauer, D. Schweinfurth, N. Deibel, B. Sarkar, H. Kelm, H.-J. Krüger, W. Frey and R. Peters, J. Am. Chem. Soc., 2012, 134, 4683–4693. G. Guilera, M. A. Newton, C. Polli, S. Pascarelli, M. Guinó and K. K. Hii, Chem. Commun., 2006, 0, 4306–4308. R. Jagadeesan, G. Velmurugan and P. Venuvanalingam, RSC Adv., 2015, 5, 80661–80667. T. W. Lyons and M. S. Sanford, Chem. Rev., 2010, 110, 1147–1169. D. C. Powers and T. Ritter, Nat. Chem., 2009, 1, 419–419. M. C. Warner and J.-E. Bäckvall, Acc. Chem. Res., 2013, 46, 2545–2555. Y. Dang, S. Qu, J. W. Nelson, H. D. Pham, Z.-X. Wang and X. Wang, J. Am. Chem. Soc., 2015, 137, 2006–2014. W. Liu, J. Zheng, Z. Liu, W. Hu, X. Wang and Y. Dang, ACS Catal., 2018, 8, 7698–7709. This journal is © The Royal Society of Chemistry 20xx J. Name., 2013, 00, 1-3 | 7 Please do not adjust margins Catalysis Science & Technology Accepted Manuscript Page 7 of 8 Open Access Article. Published on 25 March 2019. Downloaded on 3/29/2019 10:27:03 AM. This article is licensed under a Creative Commons Attribution 3.0 Unported Licence. Catalysis Science & Technology Page 8 of 8 View Article Online In situ XAS is used to reveal the evolution of palladium species during an undirected C H DOI: acetoxylation 10.1039/C8CY02430H reaction. Catalysis Science & Technology Accepted Manuscript