New Lamellar Silver Thiolate Coordination Polymers with Tunable

Dec 11, 2018 - New Lamellar Silver Thiolate Coordination Polymers with Tunable Photoluminescence Energies by Metal Substitution. Oleksandra Veselskaâ€...
0 downloads 0 Views 3MB Size
Article pubs.acs.org/IC

Cite This: Inorg. Chem. XXXX, XXX, XXX−XXX

New Lamellar Silver Thiolate Coordination Polymers with Tunable Photoluminescence Energies by Metal Substitution Oleksandra Veselska,† Caroline Dessal,† Sihem Melizi,† Nathalie Guillou,‡ Darjan Podbevšek,§ Gilles Ledoux,§ Erik Elkaim,∥ Alexandra Fateeva,⊥ and Aude Demessence*,†

Downloaded via TULANE UNIV on December 12, 2018 at 00:21:14 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.



Univ Lyon, Université Claude Bernard Lyon 1, CNRS, Institut de Recherches sur la Catalyse et l’Environnement de Lyon (IRCELYON), 69626 Villeurbanne, France ‡ Université de Versailles Saint-Quentin-en Yvelines, Université Paris-Saclay, CNRS, Institut Lavoisier de Versailles (ILV), F-78035 Versailles, France § Univ Lyon, Université Claude Bernard Lyon 1, CNRS, Institut Lumière Matière (ILM), 69626 Villeurbanne, France ∥ Synchrotron Soleil, Beamline Cristal, 91192 Gif-sur-Yvette, France ⊥ Univ Lyon, Université Claude Bernard Lyon 1, CNRS, Laboratoire des Multimatériaux et Interfaces (LMI), 69626 Villeurbanne, France S Supporting Information *

ABSTRACT: The structures of two lamellar silver thiolate coordination polymers [Ag(p-SPhCO2H)]n (1) and [Ag(pSPhCO2Me)]n (2) are described for the first time. Their inorganic part is composed of distorted Ag3S3 honeycomb networks separated by noninterpenetrated thiolate ligands. The main difference between the two compounds arises from dimeric hydrogen bonds present for the carboxylic acids. Indepth photophysical studies show that the silver thiolates exhibit multiemission properties, implying luminescence thermochromism. More interestingly, the synthesis of a heterometallic lamellar compound, [Ag 0.85 Cu 0.15 (pSPhCO2H)]n (3), allows to obtain mixed metal thiolate coordination polymers and to tune the photophysical properties with the excitation wavelengths from a green vibronic luminescence to a single red emission band.



ties.8 Among all the MOCs, a limited number of 2D materials have been isolated and structurally fully characterized. From the d10 coinage, six compounds of MOCs are reported: three with Cu(I), two with Au(I), and one with Ag(I). All these 2D MOCs are built up with para-functionalized thiophenolates, except the [Ag(SePh)] n, which is obtained with the benzeneselenolate.9 In [Cu(p-SPhR)]n (R = OH,7b CO2H8e and CO2Me8d) structures, the copper atoms have a trigonal geometry and are coordinated with three μ3-bridging sulphur atoms of the thiolate ligands to generate a lamellar Cu3S3 hexagon network with the thiolate moieties above and below the layers. In case of [Au(p-SPhR)]n (R = CO2H8c and CO2Me8b) compounds, the layers are made of chains of −S− Au−S−Au− interacting together through aurophilic interactions. In the lamellar [Ag(SePh)]n CP, the inorganic layer is made of tetragonal silver atoms and μ4-bridging selenium atoms. Although silver thiolates have been largely studied as layered silver alkanethiolates10 and their antimicrobial properties were reported,11 only one basic structural model of a

INTRODUCTION Since the discovery of graphene, intensive efforts have been directed to the synthesis of ultrathin 2D nanomaterials because they exhibit unusual chemical, electronic, and physical properties compared with their bulk counterparts.1 2D layered transition-metal chalcogenides (TMCs) appeared as a new generation of lamellar inorganic compounds with a large modulation of number of layers, composition, and structures that offer the possibility to tune their band gap energy, as well as the carrier type, leading to great interests for applications in opto/electronics, energy storage, and catalysis.2 One of the main approaches to get an isolated nanolayer of TMCs is by chemical insertion of chalcogenol molecules in a preformed TMC material in order to separate the sheets.3 The direct synthesis of layered metal organic chalcogenolates (MOCs)4 is another approach to obtain hybrid 2D materials, which combines the topology of the inorganic metal−chalcogen layers of the TMCs and a functional organic molecule.5 Recently, the anisotropic MOC coordination polymers (CPs) with d10 coinage metals have regained interest for their directional conductivity along the inorganic 1D chains6 or 2D layers7 and also for their appealing photoluminescent proper© XXXX American Chemical Society

Received: May 8, 2018

A

DOI: 10.1021/acs.inorgchem.8b01257 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry lamellar [Ag(SR)]n was proposed in the 90’s early work of Dance et al.12 No 2D silver thiolate crystallographic structure was reported so far. The layered structure of the silver benzenethiolate has been investigated to simulate the semiconductor-to-metal transition in this material which was dependent on the number of layers.13 The only heterotopic ligand-based CP structures to be reported are [Ag(2pyridinethiolate)] n 1 4 and [Ag 4 (2-mercaptobenzoate)2(H2O)2]n.15 Finally, in addition to their antibacterial activity,16 silver(I) CPs are also known for their luminescent properties, with mostly blue-green emission because of a longlived phosphorescence from ligand-centered (LC) excited states of 3ππ* character.17 Red emission, involving the metal centers, is usually observed for high nuclearity silver(I) thiolate clusters.18 Here, we report the syntheses and the first crystallographic structures of two lamellar silver thiolates, exhibiting a hexagonal Ag(I)−S layout quite similar to the 2D [Cu(SR)]n CPs. These structural similarities allowed the obtaining of a mixed-metal Ag/Cu thiolate compound with original photophysical properties. The influences of both the organic part and the metal doping on the structure and luminescence properties modulation are discussed.19



Figure 2. Structure of 2, (a) representation of the Ag3S3 network on the (ac) plane, (b) view of the distorted Ag3S3 cycle and (c) view of the lamellar structure. Blue, Ag(I); yellow, S; red, O; and gray, C. Hydrogen atoms are omitted for clarity.

The presence of short Ag(I)−Ag(I) distances of 2.936(2) Å for 1 and 2.973(3) Å for 2, are less than the sum of van der Waals atomic radii between two silver atoms (ca. 3.44 Å) and can be considered as argentophilic interactions (Table S2).20 The Ag−S distances are close in both structures: 2.456(4), 2.493(4), and 2.623(4) Å in 1 and 2.437(5), 2.489(5), and 2.627(5) Å in 2 (Figure S4 and Table S2). The S−Ag−S angles for both compounds are between 97° and 132° and the Ag−S−Ag angles for both the compounds are between 70° and 120° (Figure S4 and Table S2). This wide range of angles in the hexagonal rings explains their strong deformation and nonplanarity (Figures 1 and 2b). The organic part lies between the Ag(I)−S layers without interpenetration (Figures 1c, 2c, and S5). In 1, the presence of dimeric hydrogen bonds between the carboxylic acids is observed. The orientation of the ligands is ABBA along the c axis for 1 and ABAB along the b axis for 2, showing respectively an inversion center and a glide plane (Figure S6). The ligands are alternately positioned up and down from the inorganic layers and on each side of a layer they are arranged in a herringbone packing (Figures S7− S9). Thus, weak CH−π interactions between the aromatic rings are present with distances of 2.83(1) and 2.875(6) Å for 1 and 2.88(2) Å for 2. The topological structures of both compounds are quite similar to the corresponding copperbased polymers [Cu(p-SPhCO2H)]n/[Cu(p-SPhCO2Me)]n8d,e but completely different from the lamellar [Au(pSPhCO2H)]n/[Au(p-SPhCO2Me)]n, which exhibit aurophilic interactions between Au−S−Au chains.8b,c The SEM images of 1 and 2 show that the morphology of the crystallites can be described as thin platelets of a couple of μm length, typical of lamellar compounds (Figure S10). From the Fourier transform infrared spectroscopy (FT-IR) spectrum of 1, the −CO2 antisymmetric vibration from the carboxylic acids is observed at 1687 cm−1 and the broad ν(OH) band between 2500 and 3000 cm−1 is consistent with the presence of hydrogen bonds (Figures S11 and S12). On the FT-IR spectrum of 2, the −CO2 antisymmetric vibration of the ester group is observed at 1719 cm−1. In addition, the 1H and 13C solid-state NMR spectra of 1 and 2 (Figures S13 and S14) display the characteristic peaks of the acid and the ester with the acid proton at 13 ppm for 1 and the carbon of the methyl group at 53 ppm for 2. The carbon signal of the carboxyl function is shifted from 164 ppm for 1 to 170 ppm for 2, pointing out their purity. The C, H, and S elemental and

RESULTS AND DISCUSSION

Synthesis and Characterization. The synthesis of [Ag(pSPhCO2H)]n (1) and [Ag(p-SPhCO2Me)]n (2) is a reaction between AgNO3 and the corresponding ligand under hydrothermal conditions for 16 h at 120 °C (see the Supporting Information). The powder X-ray diffraction (PXRD) patterns of both compounds show predominant reflection characteristics of lamellar structures (Figure S1). Compounds 1 and 2 exhibit interlamellar distances of 3.39 and 3.82 nm, respectively, fitting the size of the two corresponding noninterpenetrated silver thiolate units. The structures of the lamellar compounds were obtained from PXRD (Figures S2, S3, and Table S1). Compound 1 crystallizes in the orthorhombic centrosymmetric space group Pbca (V = 1472.3(1) Å3) and 2 in the noncentrosymmetric Pca21 (V = 834.10(2) Å3). The Ag(I) atoms in 1 and 2 are in trigonal geometry and are connected to three bridging μ3-sulphur atoms from three different ligands, generating a honeycomb pattern with distorted hexagons Ag3S3 (Figures 1a and 2a).

Figure 1. Structure of 1, (a) representation of the Ag3S3 network on the (ab) plane, (b) view of the distorted Ag3S3 cycle and (c) view of the lamellar structure with the hydrogen bonds between the carboxylic acids (dotted red lines). Blue, Ag(I); yellow, S; red, O; and gray, C. Hydrogen atoms are omitted for clarity. B

DOI: 10.1021/acs.inorgchem.8b01257 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 3. 2D maps of the emission and excitation spectra of [Ag(p-SPhCO2H)]n (1) (a,b) and [Ag(p-SPhCO2Me)]n (2) (c,d) carried out in the solid state at 243 and 93 K. The intersections of white dashed lines correspond to the coordinates (λex, λem) in nm of the main peaks. Excitation and emission spectra of 1 (λex = 352 nm, λem = 522 nm) (e) and 2 (λex = 364 nm, λem = 489 nm) (f).

quantities of silver and copper precursors resulted in a CP where silver atoms are predominant. This could be due to the fact that the formation of copper(I)-thiolate is kinetically less favorable because the reduction of the Cu(II) precursor is first needed. Also, this observation tends to point out a higher affinity of thiolate function toward the softer acid silver(I) rather than copper(I) in these conditions.21 The FT-IR spectrum of 3 is similar to 1 with the presence of the antisymmetric vibrations of the −CO2 band from the carboxylic acids at 1684 cm−1 and confirms that they remain uncoordinated (Figures S20 and S20). The thermogravimetric analysis (Figure S22) shows a thermal stability of up to 400 °C in air and provides the inorganic (bulk silver and CuO) and organic contents of 42.6 and 57.4%, respectively (calculated: 40.7 and 59.3%), this small difference originates from minor amounts of bulk silver (1.5(2) wt %) as it has been seen on the PXRD pattern (Figure S18). Photophysical Properties. The solid-state UV−visible absorption spectra of 1 and 2 show a maximum of absorption at 342 nm (Figure S23). Thus, the measured optical band gap of 2.99 and 3.04 eV for 1 and 2, respectively, are slightly larger than in case of the copper analogues ([Cu(p-SPhCO2H)]n: 2.79 eV and [Cu(p-SPhCO2Me)]n: 2.68 eV) but lower to the Ag(I) nicotinic-based CPs (3.7 eV).22 Compound 3 exhibits a similar spectrum to 1 with an optical band gap of 2.97 eV (Figure S23), pointing out the absence of the effect of copper. Considering the highly emissive gold(I) analogues and the luminescence thermochromism of the copper(I) ones, the photophysical properties of 1 and 2 were carefully studied, including emission−excitation cartographies [with (λex, λem) in nm as the coordinates] and lifetime decay measurements as a function of temperature in the solid state. Unlike Cu(I)- and Au(I)-based materials, 1 and 2 are not luminescent at room temperature. The appearance of a weak emission band is observed at 273 K for 1 and at 243 K for 2. This band is centered in the low-energy (LE) range at (380, 700) and (368, 580) nm, for 1 and 2, respectively (Figure 3a,c). When the

thermogravimetric analyses confirm the chemical composition and purity of both compounds with a remaining silver content corresponding to the expected values: 41.2% (calculated: 41.3%) for 1 and 39.1% (calculated: 39.2%) for 2 (Figure S15). In addition, the compounds 1 and 2 start to decompose under air at 320 and 280 °C, respectively. The slightly better thermal stability of 1 may be explained by the presence of hydrogen bonds. A similar thermal behavior is observed in case of [Cu(p-SPhCO2H/Me)]n and [Au(p-SPhCO2H/Me)]n as well. Considering that compound 1 and the copper(I) analogue [Cu(p-SPhCO2H)]n are isostructural (Table S3 and Figure S16), attempts to obtain a heterometallic Cu(I)−Ag(I) CP were carried out in order to study the influence of the metal composition on the photophysical properties. First, a metal exchange strategy was evaluated, starting with either Ag(I) or Cu(I) CP but was ineffective. On the other hand, when equimolar quantities of silver(I) and copper(II) precursors were mixed with an excess of p-HSPhCO2H ligand in dimethylformamide at 60 °C, a heterometallic CP was successfully isolated. PXRD analysis, confirmed that the precipitate, corresponding to compound 3, was highly crystalline lamellar solid with a diffractogram close to the one of the silver(I) compound 1 (Figure S17). The first reflection of both compounds 1 and 3 is at 5.08° and there is no peak from the copper analogue at 5.18°. The Rietveld refinement of 3 (Figure S18) by using the structural model of 1 resulted in a good fit and pointed out the presence of copper with a refined chemical formula of [Ag 0.85 Cu 0.15 (pSPhCO2H)]n according to the Vegard’s law and agreeing to the occupancy factor that converges to 0.88 of silver content. The SEM images of 3 showed that the crystallites were thin platelets of 2 μm length with homogeneous size and shape distributions (Figure S19). The metal composition was confirmed by energy dispersive X-ray spectroscopy data giving a 0.87 of silver and 0.13 of copper contents for each thiolate (Table S4). Thus, the formation of 3 starting from equimolar C

DOI: 10.1021/acs.inorgchem.8b01257 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 4. Photophysical behaviors of [Ag0.85Cu0.15(p-SPhCO2H)]n (3) carried out in the solid state: (a) 2D map of the emission and excitation at 93 K, the intersections of white dashed lines correspond to the coordinates (λex, λem) in nm of the peaks; (b) emission spectra (λex = 372 nm) with the temperature; (c) emission spectra at 93 K with different excitation wavelengths (220−400 nm); (d) CIE chromaticity diagram of 1 (blue squares), [Cu(p-SPhCO2H)]n (green triangles) and 3 (black circles) showing the variation of the color with the excitation wavelength (220−400 nm) at 93 K.

The vibronic structure at HE with a progressive spacing of ∼1550 cm−1 is the characteristics of the vibrational modes of aromatic molecules involved in the intra ligand (IL) π−π* transitions. Therefore, the HE emission bands can be assigned to the IL luminescence because the free ligands themselves exhibit a single large emission band centered at 530 and 500 nm (at 93 K, λex = 360 nm) for p-HSPhCO2H and pHSPhCO2Me, respectively.8c,d Thus, the vibronic structure of the aromatic rings, present in the silver thiolate compounds and absent in the noncoordinated ligands, is favored through the heavy-atom effect of silver25 and the rigidity of the network involving different weak intramolecular interactions that generate aggregation-induced emission.26 Luminescence lifetime measurements of 1 and 2 at 93 K in solid state were performed with different excitation wavelengths to target both the HE and LE emissions (Figures S38, S39, and Table S5). The data cannot be fitted by the sum of exponential decays, but are well fitted with the sum of one fast exponential decay (t1, in the ns and μs ranges), a slow single exponential decay (t3, of a few ms), and a distribution of lifetimes in the intermediate region (t2, in the μs range) using a stretched exponential function (eq 1 in the Supporting Information).27 Even if these data point out the complexity of the systems, the shorter lifetime decay, t1, in the ns and μs ranges is consistent with both fluorescence and phosphorescence processes and can be tentatively assigned to the emission of the thiolate molecules as singlet and triplet states of IL or LL (ligand-to-ligand) charge transfers. This emission corresponds to the HE emission between 470 and 550 nm, as it has been measured for the free thiol molecules. The formation of excimers, between aromatic-based ligands, as it has been seen in silver(I)-anthracene-based compounds,

temperature is lowered, several emission peaks in the highenergy (HE) range are observed and are dependent on the excitation wavelengths (Figures 3e,f and S24−S33). Thus, at 93 K the predominant peaks in 1 are centered at (352, 484) and (352, 528) nm and at (364, 489) and (364, 526) nm for 2 (Figure 3b,d). The LE bands are centered at (368, 700) and (364, 650) nm for 1 and 2, respectively. Therefore, the main excitation wavelengths for 1 are: 324, 352, 368, and 396 nm and the emission wavelengths are 484, 528 and 700 nm and for 2 these are 324, 364, and 396 nm for the excitation and 489, 526, and 650 nm for the emission. The intensification of the excitation and the emission with the decrease of the temperature is because of the enhanced rigidity of the network, which effectively reduces the energy loss by nonradiation decay. The dependence of the emission with the excitation energy can be clearly seen in Figures S34−S37. For 1, only the HE bands are present for λex = 220−340 nm and from λex = 360 nm, the LE band appears, whereas in case of 2, the HE peaks and the LE peak at 650 nm are observed for λex = 220− 360 nm and from λex = 380 nm, the LE peak becomes predominant and is shifted to 700 nm. This excitationdependent emission that is opposite to classical photophysics has been already observed for several systems and has been tentatively explained by an anti Kasha’s rule mechanism and by a decrease of the number of excited states following the decrease of the excitation energy.23 Even if impurities have not been observed through the diverse characterizations, the presence of defects inducing site heterogeneity in the structure may also explain this excitation-wavelength dependent emission,24 as well as an imperceptible decomposition under the laser beam. D

DOI: 10.1021/acs.inorgchem.8b01257 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

3 and the monometallic analogues 1 and [Cu(p-SPhCO2H)]n. Indeed for 1, the HE bands are only present for λex = 220−340 nm and from λex = 360 nm, the LE peak appears but remains lower in intensity than the HE one (Figures S45−S46). For [Cu(p-SPhCO2H)]n, at λex = 220−340 nm the HE and LE bands are quite similar in intensity and from λex = 380 nm, the LE band decreases and HE becomes dominant (Figure S45).8e The behavior is opposite for 3, where, for λex = 220−352 nm, the HE bands are the most dominant ones and from λex = 360 nm, the LE band becomes prevalent and the HE peaks almost disappear (Figures 4c and S46). In addition, the LE peak centered at 690 nm for 3 is at intermediate energy between the LE emission of the copper and silver-based compounds at 660 and 700 nm, respectively. This original behavior further proves the successful metal substitution and the formation of Ag(I)− Cu(I) interactions, as it could not be observed for a combination of two solids. This study demonstrates that through a carefully chosen synthetic strategy, a heterometallic thiolate compound is accessible. Its photophysical response is original when compared to the monometallic analogues with the predominance of the LE band (Figure S47). Besides, when comparing the CIE chromaticity diagrams of the three compounds at 93 K with variable λex = 220−400 nm, a wider color variation, from green to red, is observed for the heterometallic structure (Figures 4d and S48).

cannot be proposed here because of the long centroid− centroid distances between the adjacent benzyl rings, 5.816(1) and 5.906(1) Å in 1 and 2, respectively.28 The longer lifetime decay, t3, consistent with phosphorescence processes and transitions through triplet states, can most probably originate from a ligand-to-metal charge transfer (LMCT).17a,c,e,29 This LMCT corresponds to the LE bands centered between 650 and 700 nm for both compounds. The intermediate lifetime, t2, reported for the distribution corresponds to the lifetime for which half of the photons have been emitted. The distribution of the lifetimes described by a stretched exponential with a β factor below 0.5 are generally associated with the possibility to energy transfer toward a distribution of nonradiative centers by dipole interactions and can be associated to the presence of structural/impurities defects in the CPs. The CIE (Commission Internationale de l’Eclairage) chromaticity diagram of 1 shows a luminescence thermochromism from yellow to green with the decrease of the temperature whereas 2 exhibits a small color variation in the yellow area (Tables S6 and S7 and Figures S40 and S41). Although, both compounds exhibit quite similar photophysical behaviors with a vibronic structure of IL/ligand-to-ligand charge transfer in the HE domain and a LMCT transition in the LE area, compound 1 shows a higher emission dependency on the excitation wavelength. This slight variation in photophysical properties observed for similar lamellar structures may be explained by the presence of hydrogen bonds in 1, implying more rigidity of the network. When the silver-based compounds 1 and 2 are compared with the Cu(I) and Au(I) analogues, a similarity arises in terms of multiemissive spectra with the presence of HE and LE bands. The main difference lies in the relative intensities of the bands and the temperature effect. Thus, [Cu(p-SPhCO2H)]n and [Au(p-SPhCO2H)]n exhibit the same vibronic structures in the HE emission and the involvement of the metals in the LE transitions.8c,e In case of [Cu(p-SPhCO2Me)]n,8d the two peaks in the HE are well separated giving an unprecedented triple emission and for [Au(p-SPhCO2Me)]n,8b the peak in the LE is so intense that the emission of the ligand is quasi absent. These observations show that the emission of these d10 MOCs is dependent on a subtle balance between the metal center, the ligand and its packing, and the whole structure that allows different charge transfers and leads to various intensity ratios between the HE and LE bands. In order to illustrate the role of metal centered nature in the photophysical properties, the emission spectra of 3 from 293 to 93 K were recorded (Figures 4 and S42). On the 2D map of the emission−excitation at 93 K of 3 (Figure 4a), three main peaks are present in the HE domain at λem = 485, 524 and 560 nm for λex = 344 nm and one peak centered in the LE area at λem = 690 nm for λex = 372 nm. Depending on the excitation wavelength, λexc = 344 and 372 nm, the HE (485 nm) and LE (690 nm) bands are dominant, respectively, whatever the temperature (Figures 4b and S42). Hence, for λex = 344 nm, the CIE chromaticity diagram gives a thermoluminochromism from orange to yellow with decreasing temperature, and for λex = 372 nm, the color of the compound is mostly red for this temperature range (Figure S43, Table S8). The lifetime decays of 3 are quite similar to 1 and 2, with long processes of 10 μs (81%) and 11 ms (19%) for the HE band and 127 μs (96%) for the LE peak, characteristics of the phosphorescence (Figure S44 and Table S5). The emission responses at variable excitation wavelengths at 93 K are quite different between the heterometallic compound



CONCLUSIONS The first structures of lamellar silver thiolates are reported. [Ag(p-SPhCO2H)]n and [Ag(p-SPhCO2Me)]n form distorted Ag3S3 honeycomb layers separated with noninterpenetrated ligands and herringbone packing. Thus, the presence of hydrogen bonds between the carboxylic acids has a small effect on the overall structure. The two compounds exhibit multiemission bands originating from the LC 3ππ* transitions in HE and from LMCT in LE. In addition, the possibility of copper substitution in the silver thiolate layers was demonstrated, leading to an original heterometallic material. More importantly, this bimetallic compound exhibits new photophysical properties different from the silver and copper parent compounds, with a major emission peak at low temperatures in the LE domain. This demonstrates the great potential of the layered MOCs as 2D materials with tunable photophysical properties. The opportunity to finely tune their luminescent response by metal substitution affords the possibility to clearly select a distinct emission signal depending on the excitation wavelength: green HE bands with λex = 344 nm and red LE peak with λex = 372 nm. This opens the path for optical properties where the judicious choice of the ligands and the metals will bring the targeted multiemissive materials. Further studies on the structures of the MOCs and the origin of the emission will have to be undertaken. Finally these new 2D MOCs may pave a new approach to isolate metal-sulfur layers for diverse applications in opto/electronics and catalysis domains. As it has been highlighted with isolated layers of 2D materials, the 2D MOCs may exhibit exotic condensed-matter phenomena that are absent in bulk.2b



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.8b01257. E

DOI: 10.1021/acs.inorgchem.8b01257 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

(4) Veselska, O.; Demessence, A. d10 Coinage Metal Organic Chalcogenolates: from Oligomers to Coordination Polymers. Coord. Chem. Rev. 2018, 355, 240. (5) (a) Fernandes, F. M.; Baradari, H.; Sanchez, C. Integrative strategies to hybrid lamellar compounds: an integration challenge. Appl. Clay Sci. 2014, 100, 2. (b) Demessence, A.; Rogez, G.; Rabu, P. Grafting of thiophenecarboxylates into magnetic transition metal hydroxide layers. Chem. Mater. 2006, 18, 3005. (6) (a) Zhang, Y.; Xia, T.; Yu, K. M.; Zhang, F.; Yang, H.; Liu, B.; An, Y.; Yin, Y.; Chen, X. Facile Synthesis of [Cu(SCH3)]n Nanowires with High Charge Mobility. ChemPlusChem 2014, 79, 559. (b) Yan, H.; Hohman, J. N.; Li, F. H.; Jia, C.; Solis-Ibarra, D.; Wu, B.; Dahl, J. E. P.; Carlson, R. M. K.; Tkachenko, B. A.; Fokin, A. A.; Schreiner, P. R.; Vailionis, A.; Kim, T. R.; Devereaux, T. P.; Shen, Z.-X.; Melosh, N. A. Hybrid metal-organic chalcogenide nanowires with electrically conductive inorganic core through diamondoid-directed assembly. Nat. Mater. 2016, 16, 349. (c) Al-Mahamad, L. L. G.; El-Zubir, O.; Smith, D. G.; Horrocks, B. R.; Houlton, A. A coordination polymer for the site-specific integration of semiconducting sequences into DNA-based materials. Nat. Commun. 2017, 8, 720. (d) Che, C.-M.; Li, C.-H.; Chui, S. S.-Y.; Roy, V. A. L.; Low, K.-H. Homoleptic Copper(I) Arylthiolates as a New Class of p-Type Charge Carriers: Structures and Charge Mobility Studies. Chem.Eur. J. 2008, 14, 2965. (7) (a) Huang, X.; Sheng, P.; Tu, Z.; Zhang, F.; Wang, J.; Geng, H.; Zou, Y.; Di, C.-a.; Yi, Y.; Sun, Y.; Xu, W.; Zhu, D. A two-dimensional p−d conjugated coordination polymer with extremely high electrical conductivity and ambipolar transport behaviour. Nat. Commun. 2015, 6, 7408. (b) Low, K.-H.; Roy, V. A. L.; Chui, S. S.-Y.; Chan, S. L.-F.; Che, C.-M. Highly conducting two-dimensional copper(I) 4hydroxythiophenolate network. Chem. Commun. 2010, 46, 7328. (8) (a) Lavenn, C.; Okhrimenko, L.; Guillou, N.; Monge, M.; Ledoux, G.; Dujardin, C.; Chiriac, R.; Fateeva, A.; Demessence, A. A luminescent double helical gold(i)-thiophenolate coordination polymer obtained by hydrothermal synthesis or by thermal solidstate amorphous-to-crystalline isomerization. J. Mater. Chem. C 2015, 3, 4115. (b) Lavenn, C.; Guillou, N.; Monge, M.; Podbevšek, D.; Ledoux, G.; Fateeva, A.; Demessence, A. Shedding light on an ultrabright photoluminescent lamellar gold thiolate coordination polymer, [Au(p-SPhCOOMe)]n. Chem. Commun. 2016, 52, 9063. (c) Veselska, O.; Okhrimenko, L.; Guillou, N.; Podbevšek, D.; Ledoux, G.; Dujardin, C.; Monge, M.; Chevrier, D. M.; Yang, R.; Zhang, P.; Fateeva, A.; Demessence, A. Intrinsic Dual-Emitting Gold Thiolate Coordination Polymer, [Au(+I)(p-SPhCO2H)]n, for Ratiometric Temperature Sensing. J. Mater. Chem. C 2017, 5, 9843. (d) Veselska, O.; Podbevšek, D.; Ledoux, G.; Fateeva, A.; Demessence, A. Intrinsic Triple-Emitting 2D Copper Thiolate Coordination Polymer as a Ratiometric Thermometer Working over 400 K Range. Chem. Commun. 2017, 53, 12225. (e) Veselska, O.; Cai, L.; Podbevšek, D.; Ledoux, G.; Guillou, N.; Pilet, G.; Fateeva, A.; Demessence, A. Structural Diversity of Coordination Polymers Based on a Heterotopic Ligand: Cu(II)-Carboxylate vs Cu(I)-Thiolate. Inorg. Chem. 2018, 57, 2736. (9) Cuthbert, H. L.; Wallbank, A. I.; Taylor, N. J.; Corrigan, J. F. Synthesis and Structural Characterization of [Cu 20 Se 4 (μ 3 SePh)12(PPh3)6] and [Ag(SePh)]. Z. Anorg. Allg. Chem. 2002, 628, 2483. (10) (a) Fijolek, H. G.; Grohal, J. R.; Sample, J. L.; Natan, M. J. A Facile Trans to Gauche Conversion in Layered Silver Butanethiolate. Inorg. Chem. 1997, 36, 622. (b) Levchenko, A. A.; Yee, C. K.; Parikh, A. N.; Navrotsky, A. Energetics of Self-Assembly and Chain Confinement in Silver Alkanethiolates: Enthalpy−Entropy Interplay. Chem. Mater. 2005, 17, 5428. (c) Voicu, R.; Badia, A.; Morin, F.; Lennox, R. B.; Ellis, T. H. Structure and Dynamics of Selectively Deuterated Self-Assembled Silvern-Octadecanethiolate Layered Materials†. Chem. Mater. 2001, 13, 2266. (d) Hu, L.; de la Rama, L. P.; Efremov, M. Y.; Anahory, Y.; Schiettekatte, F.; Allen, L. H. Synthesis and Characterization of Single-Layer Silver−Decanethiolate Lamellar Crystals. J. Am. Chem. Soc. 2011, 133, 4367.

Experimental details, chemicals, syntheses, crystallographic data, PXRD, SEM images, TGA, FT-IR, solid state NMR and UV−vis spectroscopies, and photoluminescence analyses (PDF) Accession Codes

CCDC 1839578 and 1839581 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Gilles Ledoux: 0000-0002-0867-1285 Aude Demessence: 0000-0002-8690-5489 Author Contributions

All authors have given approval to the final version of the manuscript. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The authors acknowledge SOLEIL for provision of synchrotron radiation facilities at the beamline CRISTAL (Proposal 20150723). C. Lorentz and R. Checa are warmly thanked for their help in solid state NMR and elemental analyses, respectively. We also acknowledge the CTμ (Centre Technologique des Microstructures of Lyon Univ) for providing SEM. O.V. thanks the Auvergne-Rhône-Alpes region for her PhD grant. The authors thank the Agence Nationale pour la Recherche for the financial support (MEMOL project ANR-16-JTIC-0004-01). Institut de Chimie de Lyon is also thanked for its supportive grant.



REFERENCES

(1) Gupta, A.; Sakthivel, T.; Seal, S. Recent development in 2D materials beyond graphene. Prog. Mater. Sci. 2015, 73, 44. (2) (a) Zhu, Y.; Peng, L.; Fang, Z.; Yan, C.; Zhang, X.; Yu, G. Structural Engineering of 2D Nanomaterials for Energy Storage and Catalysis. Adv. Mater. 2018, 30, 1706347. (b) Chhowalla, M.; Shin, H. S.; Eda, G.; Li, L.-J.; Loh, K. P.; Zhang, H. The chemistry of twodimensional layered transition metal dichalcogenide nanosheets. Nat. Chem. 2013, 5, 263. (3) (a) Zhang, G.; Liu, H.; Qu, J.; Li, J. Two-dimensional layered MoS2: rational design, properties and electrochemical applications. Energy Environ. Sci. 2016, 9, 1190. (b) Sun, Y.; Sun, Z.; Gao, S.; Cheng, H.; Liu, Q.; Piao, J.; Yao, T.; Wu, C.; Hu, S.; Wei, S.; Xie, Y. Fabrication of flexible and freestanding zinc chalcogenide single layers. Nat. Commun. 2012, 3, 1057. (c) Xu, Y.; Zhao, W.; Xu, R.; Shi, Y.; Zhang, B. Synthesis of ultrathin CdS nanosheets as efficient visible-light-driven water splitting photocatalysts for hydrogen evolution. Chem. Commun. 2013, 49, 9803. (d) Mak, K. F.; Shan, J. Photonics and optoelectronics of 2D semiconductor transition metal dichalcogenides. Nat. Photonics 2016, 10, 216. (e) Presolski, S.; Pumera, M. Covalent functionalization of MoS2. Mater. Today 2016, 19, 140. (f) Voiry, D.; Goswami, A.; Kappera, R.; de Carvalho Castro e Silva, C.; Kaplan, D.; Fujita, T.; Chen, M.; Asefa, T.; Chhowalla, M. Covalent functionalization of monolayered transition metal dichalcogenides by phase engineering. Nat. Chem. 2014, 7, 45. F

DOI: 10.1021/acs.inorgchem.8b01257 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

W. Y.; Tang, B. Z. Why Do Simple Molecules with ″Isolated″ Phenyl Rings Emit Visible Light? J. Am. Chem. Soc. 2017, 139, 16264. (c) Qian, H.; Cousins, M. E.; Horak, E. H.; Wakefield, A.; Liptak, M. D.; Aprahamian, I. Suppression of Kasha’s rule as a mechanism for fluorescent molecular rotors and aggregation-induced emission. Nat. Chem. 2016, 9, 83. (d) Berezin, A. S.; Vinogradova, K. A.; Krivopalov, V. P.; Nikolaenkova, E. B.; Plyusnin, V. F.; Kupryakov, A. S.; Pervukhina, N. V.; Naumov, D. Y.; Bushuev, M. B. ExcitationWavelength-Dependent Emission and Delayed Fluorescence in a Proton-Transfer System. Chem.Eur. J. 2018, 24, 12790. (24) Evariste, S.; Khalil, A. M.; Moussa, M. E.; Chan, A. K.-W.; Hong, E. Y.-H.; Wong, H.-L.; Le Guennic, B.; Calvez, G.; Costuas, K.; Yam, V. W.-W.; Lescop, C. Adaptive Coordination-Driven Supramolecular Syntheses toward New Polymetallic Cu(I) Luminescent Assemblies. J. Am. Chem. Soc. 2018, 140, 12521. (25) Kunkely, H.; Vogler, A. Can silver exert a heavy-atom effect in organic ligands? Intraligand phosphorescence of Ag(binap)NO3 with binap=2,2′-bis (diphenylphosphino)-1,1′-binaphthyl under ambient conditions. Inorg. Chem. Commun. 2006, 9, 866. (26) (a) Fermi, A.; Bergamini, G.; Roy, M.; Gingras, M.; Ceroni, P. Turn-on Phosphorescence by Metal Coordination to a Multivalent Terpyridine Ligand: A New Paradigm for Luminescent Sensors. J. Am. Chem. Soc. 2014, 136, 6395. (b) He, Z.; Ke, C.; Tang, B. Z. Journey of Aggregation-Induced Emission Research. ACS Omega 2018, 3, 3267. (27) Klafter, J.; Shlesinger, M. F. On the relationship among three theories of relaxation in disordered systems. Proc. Natl. Acad. Sci. 1986, 83, 848. (28) (a) Chen, J.; Neels, A.; Fromm, K. M. Excimer formation in crystalline and nanostructured coordination polymers. Chem. Commun. 2010, 46, 8282. (b) Chen, J.; Voutier, N.; Rajabi, J.; Crochet, A.; Bassani, D. M.; Fromm, K. M. Influence of anions and solvent molecules on the packing and emission spectra of coordination polymers based on silver ions and an anthracene derivative. CrystEngComm 2017, 19, 5106. (29) (a) Yam, V. W.-W.; Wong, K. M.-C. Luminescent metal complexes of d6, d8 and d10 transition metal centres. Chem. Commun. 2011, 47, 11579. (b) Kaeser, A.; Moudam, O.; Accorsi, G.; Séguy, I.; Navarro, J.; Belbakra, A.; Duhayon, C.; Armaroli, N.; Delavaux-Nicot, B.; Nierengarten, J.-F. Homoleptic Copper(I), Silver(I), and Gold(I) Bisphosphine Complexes. Eur. J. Inorg. Chem. 2014, 2014, 1345. (c) Li, X.-Y.; Wang, Z.; Su, H.-F.; Feng, S.; Kurmoo, M.; Tung, C.-H.; Sun, D.; Zheng, L.-S. Anion-templated nanosized silver clusters protected by mixed thiolate and diphosphine. Nanoscale 2017, 9, 3601.

(11) Nomiya, K.; Kondoh, Y.; Onoue, K.; Kasuga, N. C.; Nagano, H.; Oda, M.; Sudoh, T.; Sakuma, S. Synthesis and Characterization of Polymeric, Anionic Thiosalicylato-Ag(I) Complexes with Antimicrobial Activities. J. Inorg. Biochem. 1995, 58, 255. (12) Dance, I. G.; Fisher, K. J.; Banda, R. M. H.; Scudder, M. L. Layered structure of crystalline compounds silver thiolates (AgSR). Inorg. Chem. 1991, 30, 183. (13) Massote, D. V. P.; Mazzoni, M. S. C. Two-dimensional semiconductors: the case of silver thiolates. Appl. Phys. Lett. 2016, 109, 133104. (14) Su, W.; Hong, M.; Weng, J.; Cao, R.; Lu, S. A Semiconducting Lamella Polymer [{Ag(C5H4NS)}n] with a Graphite-Like Array of Silver(I) Ions and Its Analogue with a Layered Structure. Angew. Chem., Int. Ed. 2000, 39, 2911. (15) Sun, D.; Luo, G.-G.; Zhang, N.; Xu, Q.-J.; Jin, Y.-C.; Wei, Z.-H.; Yang, C.-F.; Lin, L.-R.; Huang, R.-B.; Zheng, L.-S. A lamella 2D silver(I) coordination polymer constructed from in situ generated 2mercaptobenzoic acid. Inorg. Chem. Commun. 2010, 13, 306. (16) (a) Morones-Ramirez, J. R.; Winkler, J. A.; Spina, C. S.; Collins, J. J. Silver enhances antibiotic activity against gram-negative bacteria. Sci. Transl. Med. 2013, 5, 190ra81. (b) Venkatesh, V.; Kumaran, M. D. B.; Saravanan, R. K.; Kalaichelvan, P. T.; Verma, S. Luminescent Silver-Purine Double Helicate: Synthesis, Self-Assembly and Antibacterial Action. ChemPlusChem 2016, 81, 1266. (c) Fromm, K. M. Silver coordination compounds with antimicrobial properties. Appl. Organometal. Chem. 2013, 27, 683. (17) (a) Igawa, S.; Hashimoto, M.; Kawata, I.; Hoshino, M.; Osawa, M. Photoluminescence Properties, Molecular Structures, and Theoretical Study of Heteroleptic Silver(I) Complexes Containing Diphosphine Ligands. Inorg. Chem. 2012, 51, 5805. (b) Hsu, C.-W.; Lin, C.-C.; Chung, M.-W.; Chi, Y.; Lee, G.-H.; Chou, P.-T.; Chang, C.-H.; Chen, P.-Y. Systematic Investigation of the Metal-Structure Photophysics Relationship of Emissive d10-Complexes of Group 11 Elements: The Prospect of Application in Organic Light Emitting Devices. J. Am. Chem. Soc. 2011, 133, 12085. (c) Moreno-Alcántar, G.; Nácar-Anaya, A.; Flores-Á lamo, M.; Torrens, H. Luminescent diphosphine fluorophenylthiolate silver(I) compounds that exhibit argentophilic interactions. New J. Chem. 2016, 40, 6577. (d) Yam, V. W.-W.; Au, V. K.-M.; Leung, S. Y.-L. Light-Emitting Self-Assembled Materials Based on d8 and d10 Transition Metal Complexes. Chem. Rev. 2015, 115, 7589. (e) Barbieri, A.; Accorsi, G.; Armaroli, N. Luminescent complexes beyond the platinum group: the d10 avenue. Chem. Commun. 2008, 2185. (18) (a) Li, G.; Lei, Z.; Wang, Q.-M. Luminescent Molecular Ag-S Nanocluster [Ag62S13(SBut)32](BF4)4. J. Am. Chem. Soc. 2010, 132, 17678. (b) Li, B.; Huang, R.-W.; Qin, J.-H.; Zang, S.-Q.; Gao, G.-G.; Hou, H.-W.; Mak, T. C. W. Thermochromic Luminescent Nest-Like Silver Thiolate Cluster. Chem.Eur. J. 2014, 20, 12416. (19) (a) Gromova, M.; Lefrançois, A.; Vaure, L.; Agnese, F.; Aldakov, D.; Maurice, A.; Djurado, D.; Lebrun, C.; de Geyer, A.; Schülli, T. U.; Pouget, S.; Reiss, P. Growth Mechanism and Surface State of CuInS2 Nanocrystals Synthesized with Dodecanethiol. J. Am. Chem. Soc. 2017, 139, 15748. (b) Bryks, W.; Smith, S. C.; Tao, A. R. Metallomesogen Templates for Shape Control of Metal Selenide Nanocrystals. Chem. Mater. 2017, 29, 3653. (20) Schmidbaur, H.; Schier, A. Argentophilic Interactions. Angew. Chem., Int. Ed. 2014, 54, 746. (21) Pearson, R. G. Hard and soft acids and bases, HSAB, part 1: Fundamental principles. J. Chem. Educ. 1968, 45, 581. (22) Xie, Y.-M.; Fan, Y.-Y.; Lin, F.-L.; Hu, T.; Liu, J.; Lu, C.-Z. Luminescent silver(I) tert -butylethynide compounds with nicotinic/ isonicotinic acid as ligands. J. Mol. Struct. 2017, 1150, 335. (23) (a) Lee, K.-M.; Cheng, W.-Y.; Chen, C.-Y.; Shyue, J.-J.; Nieh, C.-C.; Chou, C.-F.; Lee, J.-R.; Lee, Y.-Y.; Cheng, C.-Y.; Chang, S. Y.; Yang, T. C.; Cheng, M.-C.; Lin, B.-Y. Excitation-dependent visible fluorescence in decameric nanoparticles with monoacylglycerol cluster chromophores. Nat. Commun. 2013, 4, 1544. (b) Zhang, H.; Zheng, X.; Xie, N.; He, Z.; Liu, J.; Leung, N. L. C.; Niu, Y.; Huang, X.; Wong, K. S.; Kwok, R. T. K.; Sung, H. H. Y.; Williams, I. D.; Qin, A.; Lam, J. G

DOI: 10.1021/acs.inorgchem.8b01257 Inorg. Chem. XXXX, XXX, XXX−XXX