Next Article in Journal
Life Cycle Assessment (LCA) of Biochar Production from a Circular Economy Perspective
Previous Article in Journal
A Review on Pollution Treatment in Cement Industrial Areas: From Prevention Techniques to Python-Based Monitoring and Controlling Models
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Cavity Size Effect on Host-Guest Property of Tiara-like Structural Mn(SR)2n Nanoclusters Probed by NMR Spectroscopy

Shandong Sino-Japanese Center for Collaborative Research of Carbon Nanomaterials, College of Chemistry and Chemical Engineering, Instrumental Analysis Center of Qingdao University, Qingdao University, Qingdao 266071, China
*
Authors to whom correspondence should be addressed.
Processes 2022, 10(12), 2683; https://doi.org/10.3390/pr10122683
Submission received: 24 November 2022 / Revised: 8 December 2022 / Accepted: 10 December 2022 / Published: 13 December 2022
(This article belongs to the Section Chemical Processes and Systems)

Abstract

:
The lack of detect technology hinders the understanding of host-guest (H-G) chemical properties for thiolate-protected tiara-like structural nanoclusters (Mn(SR)2n). In this work, NMR spectroscopy is demonstrated as a powerful tool to probe the H-G structure of Mn(SR)2n both experimentally and theoretically. A low-field shifting and wide chemical shift (CS) signal of the H nucleus in CH2Cl2 is observed in the NMR spectrum of the mixture of CH2Cl2 and Pd8(PET)16 (PET is 2-phenylethanethiol), agreeing with the theoretical results that a deshielding area appears in the central cavity of Pd8(SR)16. All Mn(SR)2n own similar nucleus-independent chemical shift maps and deshielding cavities, which means that the H nucleus in small molecules trapped by Mn(SR)2n should have consistent low-field shifted CSs. However, such a phenomenon was only observed in the NMR spectrum of the mixed solution of Pd8(SR)16 and CH2Cl2, indicating that Pd8(SR)16 is the only one in the series of Pdn(SR)2n (n = 4~16) analogues that can capture a CH2Cl2, the H-G properties of Mn(SR)2n are highly dependent on their cavity sizes, and a guest molecule only inserts into the matching cavity of Mn(SR)2n. We anticipate that the realization of such convenient probe strategy will give a deeper understanding of the H-G properties of Mn(SR)2n.

Graphical Abstract

1. Introduction

A unique tiara-like framework endows thiolate-protected group 10 transition-metal nanoclusters (Mn(SR)2n (M is Ni, Pd, or Pt), an inorganic analogy of metallocrowns, with excellent physical and chemical properties, such as nonlinear absorption, stability, photoluminescence, catalytic performance, and host-guest (H-G) chemical properties [1,2,3,4,5,6,7,8]. Mn(SR)2n were considered potential host molecules since Ni6(SC2H5)12 was structurally determined [9]. The H-G structure of Mn(SR)2n was first experimentally determined in 2002 by the reporting of a tetrahydrofuran that was captured by Ni11(SC6H5)22 [10]. Then, several G@Mn(SR)2n (G denotes the guest molecule) were reported, including benzene and toluene molecules inserted into Ni10(SR)20 [3,11], a guest Ag+ ion accommodated in Pt6(SR)12 [12], and Pd8(SR)16 or Pt8(SR)16 encapsulating a small molecule, such as CH2Cl2, CH2Br2, CH2ClCH2Cl, CH3I, and I2 [13,14]. To the best of our knowledge, single crystal X-ray diffraction (SCXRD) is the only technique to obtain the H-G structures of G@Mn(SR)2n.
As we know, the size of the cavity in a host molecule plays a very prominent role in the formation of the H-G structure. Until now, Mn(SR)2n with various metal atoms, including Nin(SR)2n (n = 4~6, 8~12) [3,10,15,16], Pdn(SR)2n (n = 4~20) [17], and Ptn(SR)2n (n = 5~13) [1], were synthesized and isolated, but their large, high-quality single crystals, especially for the larger sized ones, were difficult to grow [18]. This is a huge obstacle to correlate their cavity sizes to H-G chemical properties using SCXRD. Therefore, it is highly desirable to develop an alternative analytical method for probing the H-G structure of Mn(SR)2n without a single crystal sample. The NMR technique was used to detect the H-G properties for a large number of materials [19,20,21,22,23,24]. For example, Hu et al. confirmed that the stereoselective, guest-driven, self-assembly of either Λ4- or Δ4- type Eu4L4 cages has been realized via chiral induction with R/S-BINOL or R/S-SPOL templates by NMR [20]. Schäfer et al. revealed the existence of a complex dynamic equilibria of oligomers that are formed by the host with bidentate guests based on diffusion NMR spectroscopy [21]. Mi et al. used solid-state NMR to ensure that the ratio of two host and guest molecules is 1:1:2 in interheteromacrocyclic hosts charge transfer crystals [23]. However, none has been reported for Mn(SR)2n thus far, which may be due to the complexity of their guest molecules, including the metal ion (Ag+), the small organic molecules, and the reductive inorganic molecule (I2). In this work with Pd8(SR)16 and CH2Cl2 as model host and guest molecules (Scheme 1), NMR spectroscopy is shown to be a highly effective approach to confirm the H-G structure of Mn(SR)2n both experimentally and theoretically. Furthermore, we use NMR spectroscopy to study the size-dependent characteristics of host Pdn(SR)2n NCs capturing guest molecules. In the future, we plan to use NMR spectroscopy for the detection of Mn(SR)2n capturing other molecules and expand their applications based on their H-G properties.

2. Materials and Methods

2.1. Materials

All reagents were commercially available and used as received. Palladium nitrate, sodium trifluoroacetate, and triethylamine were purchased from Sinopharm Chemical Reagent Co., Ltd, Shanghai, China. 2-phenylethanethiol, deuterochloroform, and trans-2-[3-(4-tert-butylphenyl)-2-methyl-2-propenylidene] malononitrile (DCTB) were purchased from Adamas, Shanghai, China.

2.2. Synthesis and Purification of Pdn(PET)2n

The synthesis of thiolate-protected palladium nanoclusters followed that of a previous report [17]. We dissolved 0.2 g Pd(NO3)2·2H2O in 12 mL acetonitrile. After stirring for 30 min, 210 μL 2-phenylethanethiol was added into the above solution. Then, the reaction mixture was stirred for another 15 min, and 0.5 mL triethylamine was rapidly added. The reaction was allowed to continue under constant stirring for 5 h. After the reaction stopped, the yellow precipitate was washed several times with methanol and water, and then collected by centrifugation. The crude product was dissolved in CH2Cl2, and the precipitate consisting of insoluble Pd-thiolates was removed. The as-obtained soluble product was further isolated via thin-layer chromatography. Each Pdn(PET)2n NC was purified by TLC for a minimum of three times.

2.3. Characterization

Matrix-assisted laser desorption ionization time-of-flight mass spectrometry (MALDI-TOF-MS) was collected by an autoflex speed TOF/TOF mass spectrometer (Bruker, Billerica, MA, USA) in reverse positive mode with DCTB used as the matrix. The 1H NMR and 2D COSY of the mixture of Pdn(PET)2n (5 ≤ n ≤ 16) and CH2Cl2 were recorded on a Bruker 400 MHz spectrometer (Billerica, MA, USA) at room temperature. Approximately 10 mg of each nanocluster was dissolved in 1 mL CDCl3 containing 0.1 μL of CH2Cl2 (approximately 1.6 × 10−3 mmol) (tetramethylsilane was used as internal standards).

2.4. Calculational Details

The geometrical structure of Pd8(SCH3)16, simplified by replacing the -CH2CO2CH3 groups of the experimentally obtained Pd8(SCH2CO2CH3)16 molecule with methyl groups, was used for the theoretical calculation. All theoretical calculations were carried out by the Density Functional Theory (DFT) method using the Gaussian 16 program (Revision B01) [25]. The selected functional was the hybrid-GGA functional B97-2 [26], which reaches rather constantly the lowest standard deviations among several common functionals [27]. The employed basis sets were the pcSseg-1 (for C, H, and S atoms), which are developed and optimized for NMR calculations [28], and the Stuttgart–Dresden double-ζ (SDD) with an effective core potential (ECP) (for Pd atoms) [29].
The nucleus-independent chemical shift (NICS) maps of the studied Pd8(SCH3)16 toroid structure were carried out by calculating the isotropic shielding value (δ(r)) and its ZZ component (δzz(r)) of 132,528 points in the space of 6.348 Å (12 Bohr) extension of the molecular coordinate [30]. These δ(r) and δzz(r) values were used to generate the maps to denote the shielded (negative) and deshielded (positive) areas of the system by using the Multiwfn program (Revision 3.6) [31].
In the calculation, an external magnetic field (Bext) of 1.0 T was applied perpendicular to the molecular plane, the induced magnetic field (Bind, in ppm units) over the space was related with the shielding tensor σ and the external magnetic field Bext, as was given elsewhere:
B ( r ) i n d = σ ( r ) B e x t
The isotropic shielding value δ(r) (in units of ppm) can be determined by the definition:
δ ( r ) = NICS = 1 3 Tr σ ( r )
where Trσ(r) represents the trace of the nuclear shielding tensor.

3. Results and Discussion

The host Pd8(SR)16 molecule was synthesized utilizing a procedure from a previous report [17]. In previous reports, a toluene molecule inserted into the Ni10S20 framework of Ni10(StBu)10(SR)10 with 2-ethylthioethanethiol, 2-(2-mercaptoethyl)pyridine or 2-aminoethanethiol as perpendicular thiols [3,11], indicated that the H-G properties of Mn(SR)2n were not dependent to their thiolate ligands. To avoid the formation of an intramolecular H-G structure by the thiolate carbon chain entering the Pd8(SR)16′s central cavity [16], 2-phenylethanethiol (PET), a thiol owning a large carbon tail (the outer diameter of phenyl (~5.3 Å) is larger than the cavity size of the Pd8S16 framework (~5.1 Å) [13]), was used as the ligand when we conducted the synthesis. The Pd8(PET)16 was synthesized and purified via a previous route [17]. Its chemical composition was identified by MALDI-TOF-MS and NMR, as shown in Figure 1 and Figure S1. The only peak centered at 3062.8 Da in the MS spectrum, and the perfect agreement between the experimental and simulated isotopic patterns indicated that the Pd8(PET)16 sample was of high purity (Figure S1). In the 1H NMR and COSY spectra (Figure 1), the three intriguing triplets at 2.6–3.2 ppm were assigned to the protons of the methylene groups (1 and 2 in Figure 1) within 2-phenylethanethiolate, and the peaks at 6.8–7.3 ppm corresponded to the phenyl groups (3–5 in Figure 1). The wide blank span at 3.4–6.8 ppm in the 1H NMR spectrum would be an ideal window to observe the CS shift of the H nuclei within a guest molecule.
Comparing four previously reported organic guest molecules of Pd8(SR)16, CH2Cl2, a cheap and easily available solvent, was optimized for the experiment. The reasons are as follows: First, its CS at 5.3 ppm [25] was in the middle of the observation window of Pd8(PET)16 and not disturbed by 2-phenylethanethiolate protected metal nanoclusters, e.g., Ni6(PET)12 [32], a small size Mn(SR)2n; and second, unlike other guest molecules disordered in the central cavity of Pd8(SR)16 [13], there was only one configuration in which one C-Cl bond was arranged vertically and another one was nearly horizontal when CH2Cl2 was encapsulated by Pd8(SR)16 [13], as shown in Scheme 1. It should be noted that CDCl3 was used as the solvent to characterize 1H NMR because it cannot insert into Pd8(SR)16 [13]; this would be an effective strategy to avoid the competition between solvent and guest molecules in tests.
To obtain the CS changes of the CH2Cl2 molecules before and after entering Pd8(PET)16, the 1H NMR spectra of pure CH2Cl2, pure Pd8(PET)16, and their mixture in CDCl3 solution were compared first. In the mixture solution, the molar ratio of Pd8(PET)16 to CH2Cl2 was 2:1, ensuring that all CH2Cl2 molecules could insert into Pd8(PET)16. The 1H NMR spectra of pure CH2Cl2, pure Pd8(PET)16, and their mixture in CDCl3 solution are compared in Figure 2 and Figure S2. The pure Pd8(PET)16 and the mixture solutions have the same CS signals of the H nuclei within Pd8(PET)16, but there were two changes in the H nuclei’s CS within CH2Cl2 between the pure CH2Cl2 and the mixture solutions. One change was the CS of CH2Cl2 in the mixture moving toward low field (0.07 ppm). As shown in Figure 2a, unlike the CS of pure CH2Cl2 at 5.30 ppm [33], there was only one single, central peak at 5.37 ppm corresponding to CH2Cl2 in the 1H NMR spectrum of the mixture. Another change was the CS signal of CH2Cl2 in the mixture solution owning a wider full width at half maximum (FWHM). As shown in Figure 2b,c, the FWHM of the mixture is approximately 0.0220 ppm, which is ~8.5 fold wider than that of pure CH2Cl2 (approximately 0.0026 ppm).
To verify the theory that the CS of H within CH2Cl2 shifting to low field was caused by CH2Cl2 inserted into Pd8(PET)16, we drew the NICS map of the host Pd8(SR)16 molecule to understand its magnetic response under an extra magnetic field. The structure of Pd8(SCH3)16 (Scheme 1), simplified from the crystal structure of Pd8(SCH2CO2CH3)16 [13], was used to carry out the calculation. After analyzing the crystal structures of Mn(SR)2n [19,20,21,34], especially Pd8(SR)16 [13,14,35], with or without guest molecules (Table 1), two main factors were taken into account when we used the simplified structure: First, the structural features of Mn(SR)2n, such as the tiara-like framework, bond length and angles, and spatial arrangements of α-CH2 (the methylene attached to the S atom), were slightly affected by their ligand species and guest molecule species; and second, for most H-G structures of G@Mn(SR)2n, the ability of the host Mn(SR)2n trapping guest molecules into their specific MnS2n toroids was independent of their thiolate ligands. So it was reasonable that we considered the nearest segments of Mn(SR)2n to the guest molecule, including the MnS2n framework and the α-CH2, when we conducted the calculation. Theoretical induced magnetic fields of 132,528 points around the Pd8(SCH3)16 were performed to investigate the isotropic magnetic response of this area. The NICS maps for Pd8(SCH3)16 are shown in Figure 3. As shown in Figure 3a, the δ(r) values in the areas nearby and outside of the molecular framework of Pd8(SCH3)16 were negative and slightly positive, indicating that these areas are shielded and slightly deshielded regions, respectively. The δ(r) values in the central cavity of the Pd8 ring were positive (dark blue), showing that the cavity is a deshielded area. From Figure 3b, the deshielded area could be further understood as a closed area (inside the bold solid black line) surrounded by the Pd8(SCH3)16 molecule. This closed, deshielded area could be characterized better by the positive values of the ZZ component of δ(r) (δzz(r) values, see Figure 3c,d). Therefore, one can infer that, once a small molecule is captured by Pd8(SCH3)16, the resonant frequency of the H nuclei within the guest molecule will move towards the low field under an external magnetic field, showing that the CS of CH2Cl2 in the mixture moving towards low field was caused by the CH2Cl2 inserting into the central cavity of Pd8(PET)16.
The wider CS signal of CH2Cl2 in the mixture solution was caused by CH2Cl2 encapsulated by Pd8(PET)16 as well. As we know, in NMR spectrum, the total observed width of the absorption line (Δν) is used to define an effective relaxation time (T) by means of the relation
Δ ν T 1
where the T is mainly determined by the spin-spin relaxation time (T2). When low concentration CH2Cl2 was dissolved in CDCl3, the CH2Cl2 could effectively produce spin-lattice relaxation due to its non-restricted thermal motion but generate a slow spin-spin relaxation process because CH2Cl2 was mainly surrounded by CDCl3; the long distance between the two H nuclei would cause a low probability of transverse relaxation. Once a CH2Cl2 was encapsulated into the cavity of the Pd8S16 framework, its spin-lattice relaxation became more difficult due to its restricted thermal motion, but its spin-spin relaxation became faster (a shorter T2) because of the fixed relative position between CH2Cl2 and Pd8(PET)16, and 144 hydrogen atoms in one Pd8(PET)16 molecule, especially 16 hydrogen atoms in the perpendicular α-CH2 of eight 2-phenylethanethiolate ligands very close to CH2Cl2. The shorter T2 would be the main factor of T for the mixture, which led to a wider absorption peak. Herein, both experimental and theoretical results show that 1H NMR spectroscopy is an effective method to detect the H-G structure of CH2Cl2@Pd8(SR)16.
Finally, the size-dependent character of Mn(SR)2n about their H-G properties, the issue which has never been studied before, was uncovered by 1H NMR spectroscopy here. Previous calculated results, which showed that several analogues of Nin(SR)2n own similar induced magnetic fields as Pd8(SR)16 [36], indicated that the induced magnetic fields of Mn(SR)2n were unrelated to their sizes and metal element species, and all Pdn(SR)2n owned a similar isotropic magnetic response as Pd8(SR)16, which meant that the CS of the H nuclei within a guest molecule would have uniform change when the small molecule inserts into the central cavity of Pdn(SR)2n. To eliminate the disturbance from ligands and treated history, the series of Pdn(PET)2n (5 ≤ n ≤ 16) analogues were synthesized and purified via the same process [17]. In the MALDI-TOF-MS spectrum of each Pdn(PET)2n (5 ≤ n ≤ 16) (Figure 4a), the single molecular ion peak in each mass spectrum indicated that every cluster sample was pure. The 1H NMR and 2D COSY spectra of CH2Cl2 mixed with each Pdn(SR)2n are shown in Figure 4b and Figures S3–S8. From the NMR data, one can easily find that the CS of CH2Cl2 in the mixture of CH2Cl2 and each Pdn(SR)2n (n = 7, 9~16) has a similar position at 5.3 ppm and width of FWHM as pure CH2Cl2, except that of the mixture of CH2Cl2 and Pd8(SR)16, indicating that only Pd8(SR)16 can capture a CH2Cl2 into its central cavity; the other sizes of Pdn(SR)2n cannot form a H-G structure with CH2Cl2. The smaller Pdn(SR)2n cannot capture CH2Cl2 perhaps due to their narrower central cavity, but the larger ones cannot capture CH2Cl2 perhaps due to the weak interaction force between the host and guest molecules that make it difficult to bind the guest molecules inserting into the larger cavity. Such results mean that a guest molecule only inserts into the matching cavity of Mn(SR)2n, and 8 is the ‘magic number’ for Pdn(SR)2n to capture CH2Cl2. The unchanged CS signal of CH2Cl2 mixed with smaller size Pdn(SR)2n, such as Pd7(SR)14, also indicates that the change in the CS signal of CH2Cl2 mixed with Pd8(SR)16 does not originate from CH2Cl2 existing in the outer deshielded area of Pd8(SR)16.

4. Conclusions

In summary, 1H NMR spectroscopy is shown to be a highly effective approach to confirm the H-G structure of Mn(SR)2n. First, the result acquired by 1H NMR spectroscopy shows that the CS of the H nuclei within CH2Cl2 shifts to low field (approximately 0.07 ppm) and widens when CH2Cl2 is mixed with Pd8(PET)16 in a CDCl3 solution. Then, the deshielded effect in the inner void of the double-crown Pd8(SR)16 was clarified via the NICS procedure, which indicated that the CS of the nucleus within a small molecule will theoretically move towards low field when the guest molecule inserts into the cluster. The faster transverse relaxation process between the hydrogen nuclei of the guest CH2Cl2 and host Pd8(PET)16 leads to a wider CS signal of the hydrogen nuclei within CH2Cl2. Both the experimental and theoretical results indicated that 1H NMR spectroscopy is a powerful tool to probe the H-G property of Mn(SR)2n. Finally, we find that Pd8(SR)16 is the only NC in the series of Pdn(SR)2n analogues that can capture a CH2Cl2 in its central cavity, showing that the H-G properties of Mn(SR)2n are highly dependent on their cavity size, and a guest molecule just inserts into the matching cavity of Mn(SR)2n. This research provides an alternative strategy to study the H-G properties of thiolate-protected tiara-like structural Mn(SR)2n nanoclusters.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/pr10122683/s1, Figure S1: the MALDI-TOF-MS spectrum of Pd8(PET)16; Figure S2: The 1H NMR spectra of pure Pd8(PET)16, the mixture of CH2Cl2+Pd8(PET)16 and pure CH2Cl2; Figure S3: 1H NMR and COSY spectra of Pd5(PET)10; Figure S4: 1H NMR and COSY spectra of Pd6(PET)12; Figure S5: 1H NMR spectrum of Pd7(PET)14 and the comparison of the 1H NMR spectra of Pd7(PET)14 and Pd6(PET)12 in the aliphatic regio; Figure S6: 1H NMR spectrum of Pd9(PET)18 and the comparison of the 1H NMR spectra of Pd9(PET)18 and Pd8(PET)16 in the aliphatic region; Figure S7: 1H NMR and COSY spectra of Pd10(PET)20; Figure S8: 1H NMR spectra of Pdn(PET)2n (11 ≤ n ≤ 16) in the range of 0~8 and 2~4.

Author Contributions

Methodology, C.Z.; software, S.G.; writing—original draft preparation, J.C.; writing—review and editing, Z.W. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the China Postdoctoral Science Foundation (2016M602093) and Natural Science Foundation of Shandong (ZR2020MB063) and the Taishan Scholar Program of Shandong Province (No. ts201511027).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Imaoka, T.; Akanuma, Y.; Haruta, N.; Tsuchiya, S.; Ishihara, K.; Okayasu, T.; Chun, W.-J.; Takahashi, M.; Yamamoto, K. Platinum clusters with precise numbers of atoms for preparative-scale catalysis. Nat. Commun. 2017, 8, 688. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Ananikov, V.P.; Orlov, N.V.; Zalesskiy, S.S.; Beletskaya, I.P.; Khrustalev, V.N.; Morokuma, K.; Musaev, D.G. Catalytic adaptive recognition of thiol (SH) and selenol (SeH) groups toward synthesis of functionalized vinyl monomers. J. Am. Chem. Soc. 2012, 134, 6637–6649. [Google Scholar] [CrossRef] [PubMed]
  3. Zhang, C.; Matsumoto, T.; Samoc, M.; Petrie, S.; Meng, S.; Corkery, T.C.; Stranger, R.; Zhang, J.; Humphrey, M.G.; Tatsumi, K. Dodecanuclear-ellipse and decanuclear-wheel nickel(II) thiolato clusters with efficient femtosecond nonlinear absorption. Angew. Chem. Int. Ed. 2010, 49, 4209–4212. [Google Scholar] [CrossRef] [PubMed]
  4. Tafen, D.N.; Kauffman, D.R.; Alfonso, D.R. Electrocatalytic oxygen evolution with pure and substituted M6(SR)12 (M = Pd, Fe, Rh) complexes. Comput. Mater. Sci. 2018, 150, 283–290. [Google Scholar] [CrossRef] [Green Version]
  5. Chen, J.; Pan, Y.; Wang, Z.; Zhao, P. The fluorescence properties of tiara like structural thiolated palladium clusters. Dalton Trans. 2017, 46, 12964–12970. [Google Scholar] [CrossRef] [Green Version]
  6. Zhu, M.; Zhou, S.; Yao, C.; Liao, L.; Wu, Z. Reduction-resistant and reduction-catalytic double-crown nickel nanoclusters. Nanoscale 2014, 6, 14195–14199. [Google Scholar] [CrossRef]
  7. Zhuang, Z.; Yang, Q.; Chen, W. One-step rapid and facile synthesis of subnanometer-sized Pd6(C12H25S)11 clusters with ultra-high catalytic activity for 4-nitrophenol reduction. ACS Sustain. Chem. Eng. 2019, 7, 2916–2923. [Google Scholar] [CrossRef]
  8. Joya, K.S.; Sinatra, L.; AbdulHalim, L.G.; Joshi, C.P.; Hedhili, M.N.; Bakr, O.M.; Hussain, I. Atomically monodisperse nickel nanoclusters as highly active electrocatalysts for water oxidation. Nanoscale 2016, 8, 9695–9703. [Google Scholar] [CrossRef] [Green Version]
  9. Woodward, P.; Dahl, L.F.; Abel, E.W.; Crosse, B.C. A new type of cyclic transition metal complex, [Ni(SC2H5)2]6. J. Am. Chem. Soc. 1965, 87, 5251–5253. [Google Scholar] [CrossRef]
  10. Ivanov, S.A.; Kozee, M.A.; Merrill, W.A.; Agarwal, S.; Dahl, L.F. Cyclo-[Ni(μ2-SPh)2]9 and cyclo-[Ni(μ2-SPh)2]11: New oligomeric types of toroidal nickel(II) thiolates containing geometrically unprecedented 9- and 11-membered ring systems. J. Chem. Soc. Dalton Trans. 2002, 22, 4105–4115. [Google Scholar] [CrossRef]
  11. Zhang, C.; Takada, S.; Kölzer, M.; Matsumoto, T.; Tatsumi, K. Nickel(II) thiolate complexes with a flexible cyclo-{Ni10S20} framework. Angew. Chem. Int. Ed. 2006, 45, 3768–3772. [Google Scholar] [CrossRef] [PubMed]
  12. Shichibu, Y.; Yoshida, K.; Konishi, K. Hexanuclear platinum(II) thiolate macrocyclic host: Charge-transfer-driven inclusion of a AgI ion guest. Inorg. Chem. 2016, 55, 9147–9149. [Google Scholar] [CrossRef] [PubMed]
  13. Yamashina, Y.; Kataoka, Y.; Ura, Y. Tiara-like octanuclear palladium(II) and platinum(II) thiolates and their inclusion complexes with dihalo- or iodoalkanes. Inorg. Chem. 2014, 53, 3558–3567. [Google Scholar] [CrossRef] [PubMed]
  14. Yamashina, Y.; Kataoka, Y.; Ura, Y. Inclusion of an iodine molecule in a tiara-like octanuclear palladium thiolate complex. Eur. J. Inorg. Chem. 2014, 2014, 4073–4078. [Google Scholar] [CrossRef]
  15. Pan, Y.; Chen, J.; Gong, S.; Wang, Z. Co-synthesis of atomically precise nickel nanoclusters and the pseudo-optical gap of Ni4(SR)8. Dalton Trans. 2018, 47, 11097–11103. [Google Scholar] [CrossRef]
  16. Dance, I.G.; Scudder, M.L.; Secomb, R. c-Ni8(SCH2COOEt)16, a receptive octagonal toroid. Inorg. Chem. 1985, 24, 1201–1208. [Google Scholar] [CrossRef]
  17. Chen, J.; Liu, L.; Weng, L.; Lin, Y.; Liao, L.; Wang, C.; Yang, J.; Wu, Z. Synthesis and properties evolution of a family of tiara-like phenylethanethiolated palladium nanoclusters. Sci. Rep. 2015, 5, 16628. [Google Scholar] [CrossRef] [Green Version]
  18. Mezei, G.; Zaleski, C.M.; Pecoraro, V.L. Structural and functional evolution of metallacrowns. Chem. Rev. 2007, 107, 4933–5003. [Google Scholar] [CrossRef]
  19. Sobiech, T.A.; Zhong, Y.; Miller, D.P.; McGrath, J.K.; Scalzo, C.T.; Redington, M.C.; Zurek, E.; Gong, B. Ultra-tight host-guest binding with exceptionally strong positive cooperativity. Angew. Chem. Int. Ed. 2022, 61, e202213467. [Google Scholar] [CrossRef]
  20. Hu, S.-J.; Guo, X.-Q.; Zhou, L.-P.; Yan, D.-N.; Cheng, P.-M.; Cai, L.-X.; Li, X.-Z.; Sun, Q.-F. Guest-driven self-assembly and chiral induction of photofunctional lanthanide tetrahedral cages. J. Am. Chem. Soc. 2022, 144, 4244–4253. [Google Scholar] [CrossRef]
  21. Schäfer, F.; Mix, A.; Cati, N.; Lamm, J.-H.; Neumann, B.; Stammler, H.-G.; Mitzel, N.W. Host-guest chemistry of a bidentate silyl-triflate bis-Lewis acid—Complex complexation behaviour unravelled by diffusion NMR spectroscopy. Dalton Trans. 2022, 51, 7164–7173. [Google Scholar] [CrossRef] [PubMed]
  22. Boles, J.E.; Bennett, C.; Baker, J.; Hilton, K.L.F.; Kotak, H.A.; Clark, E.R.; Long, Y.; White, L.J.; Lai, H.Y.; Hind, C.K.; et al. Establishing the selective phospholipid membrane coordination, permeation and lysis properties for a series of ‘druggable’ supramolecular self-associating antimicrobial amphiphiles. Chem. Sci. 2022, 13, 9761–9773. [Google Scholar] [CrossRef]
  23. Mi, Y.; Ma, J.; Liang, W.; Xiao, C.; Wu, W.; Zhou, D.; Yao, J.; Sun, W.; Sun, J.; Gao, G.; et al. Guest-binding-induced interhetero hosts charge transfer crystallization: Selective coloration of commonly used organic solvents. J. Am. Chem. Soc. 2021, 143, 1553–1561. [Google Scholar] [CrossRef] [PubMed]
  24. Wong, Y.-S.; Ng, M.; Yeung, M.C.-L.; Yam, V.W.-W. Platinum(II)-based host–guest coordination-driven supramolecular Co-assembly assisted by Pt···Pt and π–π stacking interactions: A dual-selective luminescence sensor for cations and anions. J. Am. Chem. Soc. 2021, 143, 973–982. [Google Scholar] [CrossRef] [PubMed]
  25. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 16, Revision B.01; Gaussian, Inc.: Wallingford, UK, 2016. [Google Scholar]
  26. Wilson, P.J.; Bradley, T.J.; Tozer, D.J. Hybrid exchange-correlation functional determined from thermochemical data and ab initio potentials. J. Chem. Phys. 2001, 115, 9233–9242. [Google Scholar] [CrossRef] [Green Version]
  27. Flaig, D.; Maurer, M.; Hanni, M.; Braunger, K.; Kick, L.; Thubauville, M.; Ochsenfeld, C. Benchmarking hydrogen and carbon NMR chemical shifts at HF, DFT, and MP2 levels. J. Chem. Theory Comput. 2014, 10, 572–578. [Google Scholar] [CrossRef] [PubMed]
  28. Jensen, F. Segmented contracted basis sets optimized for nuclear magnetic shielding. J. Chem. Theory Comput. 2015, 11, 132–138. [Google Scholar] [CrossRef]
  29. Andrae, D.; Haussermann, U.; Dolg, M.; Stoll, H.; Preuss, H. Energy-adjusted ab initio pseudopotentials for the second and third row transition elements. Theor. Chim. Acta 1990, 77, 123–141. [Google Scholar] [CrossRef]
  30. Chen, Z.; Wannere, C.S.; Corminboeuf, C.; Puchta, R.; Schleyer, P.V.R. Nucleus-independent chemical shifts (NICS) as an aromaticity criterion. Chem. Rev. 2005, 105, 3842. [Google Scholar] [CrossRef]
  31. Lu, T.; Chen, F. Multiwfn: A multifunctional wavefunction analyzer. J. Comput. Chem. 2012, 33, 580–592. [Google Scholar] [CrossRef]
  32. Kagalwala, H.N.; Gottlieb, E.; Li, G.; Li, T.; Jin, R.; Bern-hard, S. Photocatalytic hydrogen generation system using a nickel-thiolate hexameric cluster. Inorg. Chem. 2013, 52, 9094–9101. [Google Scholar] [CrossRef] [PubMed]
  33. Gottlieb, H.E.; Kotlyar, V.; Nudelman, A. NMR chemical shifts of common laboratory solvents as trace impurities. J. Org. Chem. 1997, 62, 7512–7515. [Google Scholar] [CrossRef] [PubMed]
  34. Tan, C.; Jin, M.; Ma, X.; Zhu, Q.; Huang, Y.; Wang, Y.; Hu, S.; Sheng, T.; Wu, X. In situ synthesis of nickel tiara-like clusters with two different thiolate bridges. Dalton Trans. 2012, 41, 8472–8476. [Google Scholar] [CrossRef]
  35. Higgins, J.D.; William Suggs, J. Preparation, structure and spectroscopic studies of the palladium mercaptides Pd8(S-nPr)16 and Pd6(S-nPr)12. Inorg. Chim. Acta 1988, 145, 247–252. [Google Scholar] [CrossRef]
  36. Muñoz-Castro, A. Bonding and magnetic response properties of several toroid structures. Insights of the role of Ni2S2 as a building block from relativistic density functional theory calculations. J. Phys. Chem. A 2011, 115, 10789–10794. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. A CH2Cl2 molecule inserts into the central cavity of Pd8(SCH3)16. Color labels: green, Pd; yellow, S; light green, Cl; grey, C; white, H.
Scheme 1. A CH2Cl2 molecule inserts into the central cavity of Pd8(SCH3)16. Color labels: green, Pd; yellow, S; light green, Cl; grey, C; white, H.
Processes 10 02683 sch001
Figure 1. 2D COSY spectra of Pd8(PET)16 nanoclusters (room temperature, CDCl3 as solvent).
Figure 1. 2D COSY spectra of Pd8(PET)16 nanoclusters (room temperature, CDCl3 as solvent).
Processes 10 02683 g001
Figure 2. (a) 1H NMR spectra, (b,c) the FWHMs of CH2Cl2; the black line is the mixture of Pd8(PET)16 and CH2Cl2, and the red line is the pure CH2Cl2 (CDCl3 as solvent).
Figure 2. (a) 1H NMR spectra, (b,c) the FWHMs of CH2Cl2; the black line is the mixture of Pd8(PET)16 and CH2Cl2, and the red line is the pure CH2Cl2 (CDCl3 as solvent).
Processes 10 02683 g002
Figure 3. NICS maps δ(r) (a,b) and its ZZ component δzz(r) (c,d) denoting shielded (negative values) and deshielded (positive values) areas in ppm for Pd8(SCH3)16 in perpendicular and parallel to Pd8 ring (xy and xz contour-plane) representations.
Figure 3. NICS maps δ(r) (a,b) and its ZZ component δzz(r) (c,d) denoting shielded (negative values) and deshielded (positive values) areas in ppm for Pd8(SCH3)16 in perpendicular and parallel to Pd8 ring (xy and xz contour-plane) representations.
Processes 10 02683 g003
Figure 4. MALDI-TOF-MS of each Pdn(PET)2n (5 ≤ n ≤ 16) (a) and 1H NMR spectra in the region from 5.25 to 5.45 ppm of the mixture of CH2Cl2 and each Pdn(PET)2n (5 ≤ n ≤ 16) (b).
Figure 4. MALDI-TOF-MS of each Pdn(PET)2n (5 ≤ n ≤ 16) (a) and 1H NMR spectra in the region from 5.25 to 5.45 ppm of the mixture of CH2Cl2 and each Pdn(PET)2n (5 ≤ n ≤ 16) (b).
Processes 10 02683 g004
Table 1. The structural information of Mn(SR)2n NCs with or without guest molecules.
Table 1. The structural information of Mn(SR)2n NCs with or without guest molecules.
SeriesH-G MoleculesBonding Length (Å)Bonding Angles (°)Ref.
M-MM-SS-α-CM-S-MS-M-SaS-M-SeM-S-α-C
Pd8(SR)16Pd8(SCH2CH2CH3)163.232.321.8488.0597.7782.27107.06[35]
Pd8(SCH2CO2CH3)163.232.321.8288.2097.6982.71107.20[13]
CH2Cl2@Pd8(SCH2CO2CH3)163.242.321.8288.4397.6682.71107.07[13]
CH2Br2@Pd8(SCH2CO2CH3)163.242.321.8188.4497.7782.62107.24[13]
(CH2Cl)2@Pd8(SCH2CO2CH3)163.242.321.8288.2997.7482.66107.00[13]
CH3I@Pd8(SCH2CO2CH3)163.252.331.8088.4697.6582.69107.12[13]
I2@Pd8(SCH2CO2CH3)163.262.331.8288.6097.7982.41106.25[14]
Pt8(SR)16Pt8(SCH2CO2CH3)163.292.321.8290.7798.7981.48107.77[13]
CH2Cl2@Pt8(SCH2CO2CH3)163.292.331.8290.3798.0482.11108.22[13]
CH2Br2@Pt8(SCH2CO2CH3)163.292.321.8390.4398.7281.56107.40[13]
(CH2Cl)2@Pt8(SCH2CO2CH3)163.292.321.8190.4198.6881.51108.18[13]
Pt6(SR)12Pt6[S-(CH2)11-CH3]123.172.321.8386.2698.4181.35108.94[12]
Ag@Pt6[S-(CH2)11-CH3]123.082.321.8383.2497.6182.39109.95[12]
Ni10-Ni10(StBu)10(SC2H5)103.152.201.8391.1896.9183.12109.67[34]
(StBu)10-CH3C6H5@Ni10(StBu)10(etet)103.212.2293.1296.6882.90[3]
(SR)10CH3C6H5@Ni10(StBu)10(pyet)103.162.211.8591.3197.4283.04109.92[3]
(0.5CH3C6H5)@Ni10(StBu)10(atet)103.172.211.8592.0696.9683.07110.02[3]
Ni10(StBu)10(mtet)103.172.211.8591.5497.4882.57110.16[11]
C6H6@Ni10(StBu)10(mtet)103.162.201.8691.5897.1082.94109.27[11]
Noted: StBu is 2-methyl-2-propanethiol, etet is 2-ethylthioethanethiolate, pyet is 2-(2-mercaptoethyl)pyridine, atet is 2-aminoethanethiol, mtet is methylthioethanethiolate.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhou, C.; Gong, S.; Chen, J.; Wang, Z. Cavity Size Effect on Host-Guest Property of Tiara-like Structural Mn(SR)2n Nanoclusters Probed by NMR Spectroscopy. Processes 2022, 10, 2683. https://doi.org/10.3390/pr10122683

AMA Style

Zhou C, Gong S, Chen J, Wang Z. Cavity Size Effect on Host-Guest Property of Tiara-like Structural Mn(SR)2n Nanoclusters Probed by NMR Spectroscopy. Processes. 2022; 10(12):2683. https://doi.org/10.3390/pr10122683

Chicago/Turabian Style

Zhou, Changlin, Shida Gong, Jishi Chen, and Zonghua Wang. 2022. "Cavity Size Effect on Host-Guest Property of Tiara-like Structural Mn(SR)2n Nanoclusters Probed by NMR Spectroscopy" Processes 10, no. 12: 2683. https://doi.org/10.3390/pr10122683

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop