Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Siloxane-linked tungsten complexes form a flexible metal–organic framework: 1-D channel structure and gas adsorption property

Ryo Nakamuraa, Koichi Nagataa, Takahiro Kawatsub, Kazuhiro Matsumoto*b, Yumiko Nakajimab, Takuji Ikedac, Shinya Takaishid, Wataru Kosakae, Hitoshi Miyasakae and Hisako Hashimoto*a
aInorganic Chemistry Laboratory, Department of Chemistry, Graduate School of Science, Tohoku University, Sendai, Miyagi 980-8578, Japan. E-mail: hisako.hashimoto.b7@tohoku.ac.jp
bInterdisciplinary Research Center for Catalytic Chemistry (IRC3), National Institute of Advanced Industrial Science and Technology (AIST), Tsukuba, Ibaraki 305-8565, Japan
cResearch Institute for Chemical Process Technology, National Institute of Advanced Industrial Science and Technology (AIST), Sendai, Miyagi 983-8551, Japan
dCoordination Chemistry Laboratory, Department of Chemistry, Graduate School of Science, Tohoku University, Sendai, Miyagi 980-8578, Japan
eInstitute for Materials Research, Tohoku University, Sendai, Miyagi 980-8577, Japan

Received 20th May 2025 , Accepted 4th July 2025

First published on 8th August 2025


Abstract

Siloxane-linked tungsten complexes were synthesised by reacting a spiro-siloxane with tungsten(methyl) complexes, and they feature covalent W–Si/W[double bond, length as m-dash]Si bonds coordinated by pyridine derivatives. Among these, [Cp*(OC)2(Ph)W[double bond, length as m-dash]Si(dmap)]2(spiro) (4) (dmap = 4-dimethylaminopyridine, spiro = spiro-siloxane) formed a porous framework with a 1-D channel structure in the crystalline state through CH–π interactions, as evaluated by single-crystal X-ray diffraction (SC-XRD) analysis. Powder X-ray diffraction (PXRD) analysis revealed that the structure of 4 reversibly changes upon the release and uptake of solvent molecules. Notably, the micro-ED measurement of the desolvated framework demonstrated a change in the space group and up to 14% contraction of its b axis, supporting the transformation of its geometrical features. Furthermore, gas adsorption measurements revealed that this framework exhibited gate-opening behavior toward gaseous molecules (5 and 4 mol mol−1 for CH4 and O2, respectively), highlighting the flexibility of the framework.


Introduction

Metal–organic frameworks (MOFs) represent a pivotal class in recent materials science, recognized for their high potential in various applications, attributable to their regulated porous structures.1–3 Since the early 2000s, increasing attention has been directed toward a subclass of flexible MOFs, referred to as soft-porous crystals (SPCs), which exhibit reversible structural transformations of the pore size and shape, triggered by external factors such as temperature, pressure or guest molecule adsorption.4–9 These properties are highly suitable for advanced gas separation, sensing and switchable catalysis. Such flexibility in MOFs can be achieved by carefully designing the coordination environment of metal nodes and selecting appropriate organic linkers. For instance, it has been reported that the incorporation of flexible aliphatic units into the linkers of frameworks induces a ‘breathing behaviour’, enabling reversible structural changes in response to the presence or absence of polar guest molecules.10,11 It has also been found that linkers bearing hydroxy, amino or amide groups, among others, enable flexible frameworks through the formation of hydrogen bonds or π–π/CH–π interactions. These frameworks exhibit reversible crystal-to-crystal transformations upon the uptake and release of molecules.12–15 More recently, the synthesis of frameworks with a catenated backbone has been reported, enabling switching between rigid and elastic states.16 These examples illustrate promising strategies for accessing various types of flexible MOFs. However, achieving high-performance MOFs that exhibit both excellent flexibility and stability remains a significant challenge.

As stable and flexible linker units, we focused our attention on siloxanes. Siloxanes, which contain Si–O–Si bonds, are widely used as silicones in various industrial applications, including silicone oil, silicone rubber and medicinal ingredients.17,18 Their versatility comes from their thermal and chemical stability, as well as their flexibility. The bond energy of the Si–O bond (549 kJ mol−1) provides stability, while the wide range of Si–O–Si bond angles (130°–180°) imparts flexibility, making them more flexible than organic compounds with a carbon-based skeleton.19 Porous materials incorporating such siloxane linkers are of great interest due to their potential achievement of flexibility. However, there have been only a few reports on the use of siloxane moieties incorporated in linkers in framework design. Davies et al. have reported the synthesis of frameworks with carboxylates containing siloxane moieties as linkers and their adsorption properties.20 Their products exhibited superior hydrophobicity and water stability, although the flexibility of the frameworks is not referred to. Cazacu et al. synthesized MOFs using a siloxane-spaced dicarboxylic acid, which became amorphous while exhibiting hydrophobicity and a slightly lower glass transition temperature, just below room temperature.21

In this study, we aimed to incorporate an inherently flexible siloxane linker to impart flexibility to the framework. By focusing on spiro-type siloxanes with conformational regulation, we also envisaged to facilitate crystallisation of the framework and to gain access to detailed structural information. Moreover, to achieve enhanced thermal stability, we opted for covalent bonds instead of coordination bonds to connect the metal units and linkers, by applying a one-pot synthesis to obtain metal–silicon bonded complexes. Also, previous studies including ours22 have demonstrated that the use of the oxidative addition/reductive elimination sequence between a labile methyl complex and hydrosilanes affords a silyl-type complex (I) and/or a silylene-type complex (II) (Scheme 1). This method is considered to produce a 16-electron intermediate (a), which facilitates the formation of silyl-type complex I via coordination of ligand L (Lewis base) to the metal. Meanwhile, silylene-type complex II is generated via a 1,2-group migration (a → b), followed by the coordination of L to the silicon atom. We anticipate that the transformability between the two forms, I and II, could also pave the way for molecular activation in future applications of this system.


image file: d5qi01162k-s1.tif
Scheme 1 Strategy for forming M–Si bonds. [M] = Cp*M(CO)n, L = labile ligand (pyridine, DMAP, etc.), R, R′ = H, alkyl, aryl, etc.

We herein report the synthesis and isolation of several di-tungsten complexes linked with spiro-siloxane. Single-crystal X-ray diffraction (SC-XRD) analysis revealed that the compound bearing 4-dimethylaminopyridine (dmap) formed a 1-D channel structure in the crystalline state. Gas adsorption measurements and micro-ED analysis under desolvation conditions demonstrated that this framework was flexible.

Results and discussion

Syntheses of siloxane-linked di-tungsten complexes

Initially, we performed the reaction of Cp*W(CO)2(py)Me (py = pyridine) with spiro-siloxane [H(Ph)Si{OSi(Me2)O}2]2Si (1)23 in a 2[thin space (1/6-em)]:[thin space (1/6-em)]1 ratio at room temperature for 2 h, which successfully afforded the siloxane-linked di-tungsten complex [Cp*(OC)2(Ph)W[double bond, length as m-dash]Si(py)]2(spiro) (2) {spiro = –(OSiMe2O)2Si(OSiMe2O)2–} in a relatively high NMR yield (75%) (Scheme 2). During the reaction, an intermediate attributed to a mono-tungsten complex with the siloxane unit was observed by 1H NMR, but it was cleanly converted into 2 in the final stage. Complex 2 was isolated in the pure form in 20% yield; the rather lower yield of 2 is due to its isomerisation to [Cp*(OC)2(py)W–Si(Ph)]2(spiro) (3) during purification. Thus, the isolated complex 2 gradually isomerised to yield 3 in 77% NMR yield, accompanied by 1,2-migration of the pyridine (from silicon to tungsten) and phenyl groups (from tungsten to silicon). Complex 3 was isolated in 34% yield based on 1 after recrystallization. Complex 2 was characterised as a base-stabilised silylene-type complex, whereas complex 3 was characterised as a silyl-type complex, based on multiple NMR measurements and the similarity of the NMR spectral data to those of the dmap analogue, [Cp*(OC)2(Ph)W[double bond, length as m-dash]Si(dmap)]2(spiro) (4), whose structure was confirmed by SC-XRD (vide infra). The di-nuclear compositions of 2 and 3 were confirmed by high-resolution mass spectrometry (HRMS), which revealed distinct molecular ion peaks.
image file: d5qi01162k-s2.tif
Scheme 2 Synthesis of 2–5 and pKa of [H+]base.27

In the 29Si{1H} NMR spectrum, complex 2 exhibited a signal of the silicon atom bonded to the tungsten atom at 27.0 ppm, with a satellite of tungsten (1JWSi = 131 Hz, 183W, nuclear spin = 1/2) (Table 1). This signal is considerably shifted downfield compared to that of the precursor spiro-siloxane 1 (−47.1 ppm), consistent with the formation of W–Si bonds. The satellite coupling constant 1JWSi also provides useful information about the bonding nature; the spin–spin coupling constant J reflects the s character of the nuclei involved in the bond.24 Thus, the 1JWSi value (131 Hz) of 2 is close to that of the previously reported complex [Cp*(OC)2(Me)W[double bond, length as m-dash]Si(OMe)2(dmap)] (A) (1JWSi = 148 Hz),25 suggesting the considerable sp2 character of the Si atom involved in the W–Si bonds in 2. In contrast, the 1JWSi value (53.8 Hz) of 3 at 21.5 ppm is much smaller than that of 2 and comparable to that of the related silyl complex [Cp*(OC)2(dmap)W–SiMe2Ph] (B) (1JWSi = 31 Hz),26 suggesting the sp3-character of the Si atom bonded to tungsten in 3. In addition, this type of isomerization from a silylene-type complex to a silyl complex has been previously found by our group; [Cp*(OC)2(Ph)W[double bond, length as m-dash]SiMe2(dmap)] (C) was converted into B via 1,2-migration of the phenyl group.26

Table 1 Selected 29Si{1H} NMR data of complexes 1–5, A and B (SiH or SiW signals)
Complex 29Si{1H} NMRa (ppm) 1JWSi (Hz)
a In C6D6.b In CD2Cl2.c From ref. 25.d From ref. 26.
[H(Ph)Si{OSi(Me2)O}2]2Si (1) −47.1 (SiH)
[Cp*(OC)2(Ph)W[double bond, length as m-dash]Si(py)]2(spiro) (2) 27.0 131
[Cp*(OC)2(py)W–Si(Ph)]2(spiro) (3) 21.5 53.8
[Cp*(OC)2(Ph)W[double bond, length as m-dash]Si(dmap)]2(spiro) (4) 22.8b 121
[Cp*(OC)2(pyBr)W–Si(Ph)]2(spiro) (5) 21.1 53.3
[Cp*(OC)2(Me)W[double bond, length as m-dash]Si(OMe)2(dmap)] (A)c 56.6 148
[Cp*(OC)2(dmap)W–SiMe2Ph] (B)d 15.8 31


Next, we envisaged to obtain more stable products by replacing the coordinated pyridine with DMAP, which has more basicity and stronger coordination ability than pyridine {pKa of [H+]base: 5.25 (H+·py) < 9.7 (H+·DMAP)}.27 As expected, the addition of DMAP to complex 2 resulted in a facile exchange of pyridine with DMAP to produce [Cp*(OC)2(Ph)W[double bond, length as m-dash]Si(dmap)]2-(spiro) (4) (Scheme 2). The pyridine/DMAP exchange also proceeded from 3 to give 4, accompanied by 1,2-Ph migration. Consequently, 4 was more directly obtained by the reaction of the dmap-methyl complex Cp*W(CO)2(dmap)Me26 and spiro-siloxane 1, from which 4 was isolated in the highest yield (84%). The 29Si{1H} NMR of 4 exhibits a signal at 22.8 ppm with a relatively large 1JWSi coupling of 121 Hz, suggesting that 4 is a base-stabilised silylene complex (Table 1); this value is larger than that of the silyl complex 3 (1JWSi = 53.8 Hz) but is close to those of silylene-type complexes A (1JWSi = 148 Hz) and 2 (1JWSi = 131 Hz). On the other hand, we also tested the reaction of 1 with a more labile methyl complex with a much weaker coordination ligand {pyBr = 3-bromopyridine, pKa of [H+]base: 2.84 (H+·pyBr) < 5.25 (H+·py)},27a as a starting precursor. Thus, the reaction of Cp*W(CO)2(pyBr)Me with 1 gave solely silyl complex 5 in a much shorter reaction time (compared with the formation of 3) (Scheme 2). The 29Si{1H} NMR signal data (21.1 ppm, 1JWSi = 53.3 Hz) agree with the observation that 5 is a silyl-type complex (Table 1): there was no observation of a silylene-type complex. Complex 5 was efficiently converted into 4 by the reaction with DMAP (90% NMR yield) due to its facile substitution ability.

Solid-state molecular structure of the tungsten di-nuclear complex 4

SC-XRD analysis clearly revealed that 4 is a di-nuclear complex bridged by a spiro-siloxane moiety (Fig. 1). Each tungsten centre of 4 adopts a four-legged piano-stool geometry, with the nitrogen atom N1 of the dmap being coordinated to the silicon atom Si1. The W–Si bond lengths [2.465(5), 2.470(6) Å] are within the range of those of base-stabilised tungsten(silylene) complexes (2.45–2.51 Å).28 The Si–N bond lengths [1.880(13), 1.86(2) Å] are longer than the usual Si–N covalent bond lengths (1.70–1.76 Å),29 suggesting the coordination character of these bonds. These lengths are slightly shorter than those of the reported nitrogen-coordinated silylene complexes [1.908(2)–2.007(9) Å],30 implying rather stronger coordination of dmap in 4. The sums of the three bond angles around the Si atoms [344.4(7)°, 343.7(8)°], excluding the Si–N bonds, are intermediate between those of the ideal tetrahedral geometry (329°) for sp3-Si and the trigonal planar geometry (360°) for sp2-Si. All these features are characteristics of a base-stabilised silylene complex and indicate the relatively strong coordination of dmap to the silylene moiety.
image file: d5qi01162k-f1.tif
Fig. 1 Crystal structure of [Cp*(OC)2(Ph)W[double bond, length as m-dash]Si(dmap)]2(spiro) (4). Selected interatomic distances (Å): W(1)–Si(1) 2.465(5), Si(1)–N(1) 1.880(13), W(2)–Si(2) 2.470(6), Si(2)–N(2) 1.86(2). Hydrogen atoms and crystal solvents are omitted for clarity.

Packing structure of complex 4

The packing structures of complex 4 are illustrated in Fig. 2a–c, along the a-, b- and c-axis directions. It is clearly observed from these figures that the complex forms a porous 1-D channel structure along the c-axis direction. The pore size is about 9 Å × 18 Å (Fig. 2d and e). The pores contain CH2Cl2 and THF as the crystal solvents, but their positions were not determined due to their random arrangement. The view focusing on the area around dmap ligands (Fig. 3) shows the existence of two types of CH–π intermolecular interactions in the crystalline state: (i) CH–π(py) interaction between the hydrogen atoms of the (dimethyl)amino groups of dmap ligands and the carbon atoms at the 3-position of the six-membered rings, (ii) CH–π(CO) interaction between the hydrogen atoms of the (dimethyl)amino groups and the carbon atoms of the CO ligands. The distances between these atoms are (i) 2.86 Å and (ii) 2.63 Å, respectively, which are within the range of the CH–π interaction length (2.6–3.0 Å).31 These multiple interactions appear to play a role in shaping the framework.
image file: d5qi01162k-f2.tif
Fig. 2 Packing structure of [Cp*(OC)2(Ph)W[double bond, length as m-dash]Si(dmap)]2(spiro) (4), along the (a) a-, (b) b- and (c) c-axis: atomic codes W, light blue; C, grey; N, purple; O, red; and Si, light yellow. Hydrogen atoms and crystal solvents are omitted for clarity. (d) View of the 1-D channel in the crystal phase of 4. (e) Enlarged view of the pore along the c-axis.

image file: d5qi01162k-f3.tif
Fig. 3 CH–π interaction in the crystal of 4 around the dmap ligands: atomic codes W, light blue; C, grey; N, purple; O, red; Si, light yellow; and H, white. Blue line: CH–π(py), green line: CH–π(CO) interaction.

Uptake of solvent molecules by complex 4

The presence of the 1-D channel structure in 4 encouraged us to investigate the adsorption property of 4 in the solid state.12,16 The powder X-ray diffraction (PXRD) pattern of the samples A–E and the simulation pattern I obtained from the SC-XRD analysis of 4 are shown in Fig. 4 (see the ESI, Fig. S2). The PXRD pattern II of sample A, which is a crystal of the mother liquor, is consistent with simulation I. The PXRD pattern III of sample B, which is a powder of 4 obtained after drying under reduced pressure, differs significantly from simulation I. This observation suggests that 4 was transformed into a different framework from that of the single-crystal structure upon drying.
image file: d5qi01162k-f4.tif
Fig. 4 Powder X-ray diffraction patterns of 4. I: the simulation pattern obtained from the SC-XRD analysis; II: the PXRD pattern of sample A – crystals in mother liq.; III: sample B – dried under vacuum; IV: sample C – dried under Ar; V: sample D – immersed in n-hexane for 19 days at r.t.; and VI: sample E – immersed in n-hexane for 47 days at r.t. (2θ = 3–30°).

Pattern III is essentially identical to pattern IV of sample C, which is crystals dried slowly under Ar, suggesting that the drying speed has no impact on this structural transformation. Moreover, this characteristic implies the ease with which the crystalline solvent can be desolvated and activated, making it highly suitable for application as a gas adsorption material. In pattern V of sample D, which consisted of dried powder B immersed in n-hexane at room temperature, superimposed features of the patterns of samples A and B were observed. These results demonstrate that immersing the dried powder of complex 4 (sample B) in n-hexane enables the solvents to penetrate the pores, gradually reconstructing the original 1-D channel structure. This process is identified as a solid-to-solid transition, rather than recrystallisation, due to the insolubility of 4 in n-hexane. Furthermore, a comparison between pattern VI of sample E, which is dried powder B immersed in n-hexane at room temperature for 47 days, and pattern V of sample D, immersed for 19 days, clearly reveals that the component corresponding to sample A increased in sample E. This result indicates that extended immersion in n-hexane leads to an increased proportion of 1-D channel structures.

Based on these observations from the PXRD patterns of samples A, B, D and E, it is evident that 4 undergoes a reversible framework transformation upon the release and uptake of solvent molecules (n-hexane). In addition, attempts were made to perform SC-XRD measurement on the dried structure; however, the detailed structure could not be determined due to its small crystalline size.

Micro-ED measurements of dried complex 4

To gain further insight into the dried structure of complex 4 (dried-4), we conducted micro-ED analysis, which revealed the lattice parameters of dried-4, although the molecular structure of dried-4 was not determined. Surprisingly, upon desolvation, the space group transformed from P21/c to P[1 with combining macron] (or P1), accompanied by a reduction in the b-axis unit cell from approximately 33 Å to 28 Å, representing a 14% decrease (Table 2). Along with the reduction of the b-axis, the cell volume decreased from approximately 8128 Å3 to 7056 Å3, representing 13% shrinkage. It is important to note that using the lattice constants obtained from micro-ED analysis, the observed PXRD pattern was well fitted by the Le Bail method (Fig. S6), indicating that the obtained powder is a single-phase. The slight difference between the lattice constants refined by the Le Bail analysis and those obtained by micro-ED may be due to the difference in measurement conditions; micro-ED measurements were carried out under vacuum conditions, while the PXRD was performed under atmospheric pressure.
Table 2 Unit cell parameters of complex 4 (SC-XRD) and dried-4 (micro-ED)
  4 Dried-4
Crystal system Monoclinic Triclinic
Space group P21/c P[1 with combining macron] (or P1)
a [Å] 26.0217(13) 26.737(11)
b [Å] 32.6729(15) 28.087(8)
c [Å] 9.5848(4) 9.577(3)
α [°] 90 95.94(3)
β [°] 94.109(4) 91.77(4)
γ [°] 90 99.01(3)
V3] 8128.1(6) 7056(4)


Gas adsorption measurements of complex 4

PXRD analysis demonstrated the capability of the framework to incorporate molecules into its pores, suggesting its potential application in gas adsorption. Accordingly, gas adsorption measurements were conducted on powder samples of complex 4 for CO2 (195 K), CH4 (112 K), O2 (90 K), N2 (77 K or 120 K), Ar (87 K) and H2 (77 K). As shown in the adsorption isotherm of CO2 gas in Fig. 5, two molecules of CO2 were adsorbed per formula unit of complex 4 (equivalent to 27.4 cm3 g−1 at standard temperature and pressure) and the adsorption isotherm illustrated a slightly stepwise adsorption behavior. However, the sorption process did not induce any substantial structural changes in the framework, as evidenced by the in situ PXRD experiments conducted under a CO2 atmosphere at 195 K (Fig. S8). In contrast, the adsorption isotherms for CH4 and O2 exhibited distinct gate-opening behavior during the adsorption of these gases (Fig. 6). Gate-opening adsorption refers to the behavior in which gas molecules are absorbed, as the framework of the adsorbent undergoes a structural transition at a specific threshold gas pressure. This phenomenon can be attributed to the structural flexibility of the framework in response to guest stimuli.1b,32 At 100 kPa, the CH4 uptake at 112 K reached 5 mol mol−1 (69.7 cm3 g−1 at STP) and the adsorption isotherm exhibited an abrupt increase at 77 kPa, along with pronounced hysteresis. In the O2 isotherm at 90 K, the adsorption capacity was determined to be 4 mol mol−1 (55.9 cm3 g−1 at STP), showing a gradual increase at 60–70 kPa and exhibiting pronounced hysteresis as well. In contrast, the adsorption of other gaseous molecules (N2, Ar and H2) was not observed (Fig. S7). The minimal N2 adsorption at 77 K can be attributed to the kinetic constraints, specifically the rigidity of the framework at low temperatures. This is supported by the slightly higher N2 adsorption observed at 120 K (3.02 cm3 g−1 at STP) than that at 77 K (1.73 cm3 g−1 at STP). Similarly, the absence of Ar (87 K) and H2 (77 K) uptake can be attributed to the kinetic constraints. Interestingly, however, clear adsorption behavior was observed for O2 gas, despite its physical and electrical properties being similar to those of Ar. This phenomenon may be ascribed to the sensitivity of the framework to subtle variations in quadrupole moment or adsorption temperature, although the specific mechanisms remain unclear (Table S4). Additionally, the ability of the framework to adsorb CH4 suggests that its selectivity may also be influenced by the polarizability of guest molecules.
image file: d5qi01162k-f5.tif
Fig. 5 Adsorption isotherm of CO2 at 195 K. The solid and open circles in blue represent adsorption and desorption, respectively.

image file: d5qi01162k-f6.tif
Fig. 6 Adsorption isotherms of CH4 (112 K) and O2 (90 K). The solid-open square and triangle represent the adsorption–desorption of CH4 and O2, respectively.

In general, the solid phase typically exists in its most stable conformation, and any conformational changes are energetically unfavorable.33 Therefore, when gate-opening adsorption takes place, as observed in the present case, stabilisation must occur to offset the destabilisation caused by the conformational changes induced during adsorption. Hence, host–guest interaction that stabilises the guest-adsorbed phase is essential. The adsorption of CH4 and O2 is believed to involve interactions between the framework and these gases (CH4 or O2), arising from the polarisation of electric charge distribution or dielectric effects. These interactions likely drive the observed adsorption behavior. Whether this dynamic behavior stems from the introduction of the siloxane remains to be investigated. Our future studies will include determining the crystal structure of the dried phase, conducting spectroscopic measurements (IR and Raman) under gas atmospheres, and performing control experiments using a porous material lacking siloxane units. Additionally, it should be noted that this siloxane-based porous material has potential applications in gas separation and sensing due to its high adsorption selectivity for O2 and CH4 over other gases (N2, Ar and H2). Future research will also focus on adsorption measurements using gas mixtures and breakthrough curve analyses.

Conclusions

In summary, this study aimed to synthesise complexes linked by siloxanes to achieve flexible structures. Di-nuclear complexes 2–5 linked with siloxane were successfully synthesised, and their structures were elucidated by multiple spectroscopic techniques, along with SC-XRD analysis of 4. The SC-XRD study confirmed that complex 4 is a di-nuclear complex featuring dmap-coordinated W[double bond, length as m-dash]Si bonds, forming a porous framework with a 1-D channel structure in its crystalline state through intermolecular CH–π interactions. The PXRD analyses reveal that the dried powder of 4 reconstructs the 1-D channel structure upon immersion in n-hexane. Moreover, gas adsorption measurements revealed that the framework adsorbs two molecules of CO2, four molecules of O2, and five molecules of CH4 per formula unit of 4. Notably, only in the cases of CH4 and O2, does this compound exhibit sorption phenomena characteristic of gate-type sorption behavior, highlighting the flexibility of the framework. This is further supported by micro-ED analysis, which demonstrated changes in the lattice parameters upon desolvation. This phenomenon is thought to result from the introduction of a flexible siloxane into the framework. Further investigations on the effect of the siloxane unit, adsorption behavior, and properties of other analogous complexes are in progress.

Author contributions

Conceptualisation, methodology, K. M., Y. N. and H. H.; synthesis, R. N., T. K.; analysis, investigation, validation, R. N., K. N., T. I., S. T., W. K., H. M.; resources, visualization, funding acquisition, K. N. and H. H.; original draft preparation, R. N.; review and editing, supervision, project administration, H. H. All authors have read and agreed to the published version of the manuscript.

Conflicts of interest

There are no conflicts to declare.

Data availability

The data supporting this article have been included in the ESI. Synthetic and experimental procedures, characterisation data, details of X-ray crystallographic analysis, powder X-ray diffraction, micro-ED, thermogravimetric analysis, Le Bail analysis, gas sorption measurement and NMR spectra. The crystallographic data for the X-ray crystallographic coordinates of structures reported in this study have been deposited at the Cambridge Crystallographic Data Centre (CCDC) under deposition number CCDC 2445984.

Acknowledgements

This work was supported by JSPS KAKENHI grants JP_ 15K05444, 22H02088, 23K23356 (HH), and 21K14638 (KN) from the Japan Society for the Promotion of Science (JSPS), and by JST SPRING, Grant Number JPMJSP2114 (RN). This work was supported by JST-FOREST (Fusion Oriented Research for Disruptive Science) (YN). We thank Mr H. Sato (Rigaku Corporation, Tokyo, Japan) for his support in XRD study and are also grateful to the Research and Analytical Center for Giant Molecules (Tohoku University) for mass and elemental analyses. The micro-ED analysis was supported by the Nanomaterial Evaluation and Prototyping Platform (NEPP) at AIST Tohoku.

References

  1. (a) S. Kitagawa and M. Kondo, Functional Micropore Chemistry of Crystalline Metal Complex-Assembled Compounds, Bull. Chem. Soc. Jpn., 1998, 71, 1739–1753 CrossRef CAS; (b) S. Kitagawa, R. Kitaura and S. Noro, Functional Porous Coordination Polymers, Angew. Chem., Int. Ed., 2004, 43, 2334–2375 CrossRef CAS PubMed; (c) S. Kitagawa and R. Matsuda, Chemistry of coordination space of porous coordination polymers, Coord. Chem. Rev., 2007, 251, 2490–2509 CrossRef CAS.
  2. J.-R. Li, J. Sculley and H.-C. Zhou, Metal–Organic Frameworks for Separations, Chem. Rev., 2012, 112, 869–932 CrossRef CAS PubMed.
  3. H. Furukawa, K. E. Cordova, M. O'Keeffe and O. M. Yaghi, The Chemistry and Applications of Metal-Organic Frameworks, Science, 2013, 341, 1230444 CrossRef PubMed.
  4. G. Ferey and C. Serre, Large breathing effects in three-dimensional porous hybrid matter: facts, analyses, rules and consequences, Chem. Soc. Rev., 2009, 38, 1380–1399 RSC.
  5. S. Horike, S. Shimomura and S. Kitagawa, Soft porous crystals, Nat. Chem., 2009, 1, 695–704 CrossRef CAS PubMed.
  6. M. Li, D. Li, M. O'Keeffe and O. M. Yaghi, Topological Analysis of Metal−Organic Frameworks with Polytopic Linkers and/or Multiple Building Units and the Minimal Transitivity Principle, Chem. Rev., 2014, 114, 1343–1370 CrossRef CAS PubMed.
  7. A. Schneemann, V. Bon, I. Schwedler, I. Senkovska, S. Kaskel and R. A. Fischer, Flexible metal–organic frameworks, Chem. Soc. Rev., 2014, 43, 6062–6096 RSC.
  8. L. Sarkisov, R. L. Martin, M. Haranczyk and B. Smit, On the Flexibility of Metal–Organic Frameworks, J. Am. Chem. Soc., 2014, 136, 2228–2231 CrossRef CAS PubMed.
  9. R. Pallach, J. Keupp, K. Terlinden, L. F. Beyme, M. Kloß, A. Machalica, J. Kotschy, S. K. Vasa, P. A. Chater, C. Sternemann, M. T. Wharmby, R. Linser, R. Schmid and S. Henke, Frustrated flexibility in metal-organic frameworks, Nat. Commun., 2021, 12, 4097 CrossRef CAS PubMed.
  10. S. Henke, D. C. F. Wieland, M. Meilikhov, M. Paulus, C. Sternemann, K. Yusenko and R. A. Fischer, Multiple phase-transitions upon selective CO2 adsorption in an alkyl ether functionalized metal–organic framework—an in situ X-ray diffraction study, CrystEngComm, 2011, 13, 6399–6404 RSC.
  11. H. Reinsch, R. S. Pillai, R. Siegel, J. Senker, A. Lieb, G. Maurin and N. Stock, Structure and properties of Al-MIL-53-ADP, a breathing MOF based on the aliphatic linker molecule adipic acid, Dalton Trans., 2016, 45, 4179–4186 RSC.
  12. K. Uemura, S. Kitagawa, M. Kondo, K. Fukui, R. Kitaura, H.-C. Chang and T. Mizutani, Novel Flexible Frameworks of Porous Cobalt(II) Coordination Polymers That Show Selective Guest Adsorption Based on the Switching of Hydrogen-Bond Pairs of Amide Groups, Chem. – Eur. J., 2002, 8, 3587–3600 CrossRef.
  13. N. Dissem, M. Essalhi, N. Ferhi, A. Abidi, T. Maris and A. Duong, Flexible and porous 2D layered structures based on mixed-linker metal–organic frameworks for gas sorption studies, Dalton Trans., 2021, 50, 8727–8735 RSC.
  14. R. Kitaura, K. Seki, G. Akiyama and S. Kitagawa, Porous Coordination-Polymer Crystals with Gated Channels Specific for Supercritical Gases, Angew. Chem., Int. Ed., 2003, 42, 428–431 CrossRef CAS PubMed.
  15. K. Roztocki, W. Gromelska, F. Formalik, A. Giordana, L. Andreo, G. Mahmoudi, V. Bon, S. Kaskel, L. J. Barbour, A. Janiak and E. Priola, Shape-Memory Effect Triggered by (−( Interactions in a Flexible Terpyridine Metal–Organic Framework, ACS Mater. Lett., 2023, 5, 1256–1260 CrossRef CAS PubMed.
  16. W. Meng, S. Kondo, T. Ito, K. Komatsu, J. Pirillo, Y. Hijikata, Y. Ikuhara, T. Aida and H. Sato, An elastic metal–organic crystal with a densely catenated backbone, Nature, 2021, 598, 298–303 CrossRef CAS PubMed.
  17. (a) E. G. Rochow, Silicon and Silicones, Springer-Verlag Berlin Heidelberg New York London Paris Tokyo, 1987 CrossRef; (b) J. Martín-Gil, F. J. Martín-Gil, A. I. De Andrés-Santos, M. C. Ramos-Sànchez, M. T. Barrio-Arredondo and N. Chebib-Abuchalà, Thermal behaviour of medical grade silicone oils, J. Anal. Appl. Pyrolysis, 1997, 42, 151–158 CrossRef; (c) H. Miyahara, A. Nakajima, J. Wada and S. Yanabu, Breakdown Characteristics of Combined Insulation in Silicone Oil for Electric Power Apparatus, 2006 IEEE 8th International Conference on Properties and applications of Dielectric Materials, 661–664; (d) R. H. Doremus, Viscosity of silica, J. Appl. Phys., 2002, 92, 7619–7629 CrossRef CAS.
  18. (a) M. Unno, A. Suto and T. Matsumoto, Laddersiloxanes—silsesquioxanes with defined ladder structure, Russ. Chem. Rev., 2013, 82, 289–302 CrossRef; (b) Y. Liu, T. Chaiprasert, A. Ouali and M. Unno, Well-defined cyclic silanol derivatives, Dalton Trans., 2022, 51, 4227–4245 RSC.
  19. (a) M. A. Brook, Silicon in Organic, Organometallic, and Polymer Chemistry, John Wiley & Sons, New York, 2000 Search PubMed; (b) F. Weinhold and R. West, The Nature of the Silicon–Oxygen Bond, Organometallics, 2011, 30, 5815–5824 CrossRef CAS.
  20. (a) L. C. Delmas, A. J. P. White, D. Pugh, A. Evans, M. A. Isbell, J. Y. Y. Heng, P. D. Lickiss and R. P. Davies, Stable metal–organic frameworks with low water affinity built from methyl-siloxane linkers, Chem. Commun., 2020, 56, 7905–7908 RSC; (b) L. C. Delmas, P. N. Horton, A. J. P. White, S. J. Coles, P. D. Lickiss and R. P. Davies, Studies on the structural diversity of MOFs containing octahedral siloxane-backboned connectors, Polyhedron, 2019, 157, 25–32 CrossRef CAS; (c) L. C. Delmas, P. N. Horton, A. J. P. White, S. J. Coles, P. D. Lickiss and R. P. Davies, Siloxane-based linkers in the construction of hydrogen bonded assemblies and porous 3D MOFs, Chem. Commun., 2017, 53, 12524–12527 RSC.
  21. M. Cazacu, G.-O. Turcan-Trofin, A. Vlad, A. Bele, S. Shova, A. Nicolescu and A. Bargan, Hydrophobic, amorphous metal–organic network readily prepared by complexing the aluminum ion with a siloxane spaced dicarboxylic acid in aqueous medium, J. Appl. Polym. Sci., 2019, 136, 47144 CrossRef.
  22. (a) H. Sakaba, M. Tsukamoto, T. Hirata, C. Kabuto and H. Horino, Synthesis, Structure, and Silylene Exchange Reaction of Base-Stabilized Hydrido(silylene)tungsten Complexes and Rearrangement of Hydrosilyl(pyridine)tungsten Complexes to the Base-Stabilized Hydrido(silylene) Complexes via 1,2-Hydrogen Migration, J. Am. Chem. Soc., 2000, 122, 11511–11512 CrossRef CAS; (b) K. Ueno, S. Asami, N. Watanabe and H. Ogino, Synthesis of Self-Stabilized and Donor-Free Silyl(silylene)tungsten Complexes, Organometallics, 2002, 21, 1326–1328 CrossRef CAS; (c) H. Tobita, A. Matsuda, H. Hashimoto, K. Ueno and H. Ogino, Direct Evidence for Extremely Facile 1,2- and 1,3-Group Migrations in an FeSi2 System, Angew. Chem., Int. Ed., 2004, 43, 221–224 CrossRef CAS PubMed.
  23. T. Kawatsu, K. Fuchise, K. Takeuchi, J.-C. Choi, K. Sato and K. Matsumoto, Well-defined hydrogen and organofunctional polysiloxanes with spiro-fused siloxane backbones, Polym. Chem., 2021, 12, 2222–2227 RSC.
  24. (a) N. F. Ramsey, Electron Coupled Interactions between Nuclear Spins in Molecules, Phys. Rev., 1953, 91, 303–307 CrossRef CAS; (b) H. McConnell, Molecular Orbital Approximation to Electron–Spin Coupled Nuclear Spin–Spin Interactions in Molecules, J. Chem. Phys., 1955, 27, 760 CrossRef; (c) H. M. McConnell, Molecular Orbital Approximation to Electron Coupled Interaction between Nuclear Spins, J. Chem. Phys., 1956, 24, 460–467 CrossRef CAS; (d) F. H. Köhler, H. J. Kalder and E. O. Fisher, Die Metall–Kohlenstoff-Bindung in Carben- und Carbin-Komplexen. Aussagen Der 183W–13C-Kopplungen, J. Organomet. Chem., 1976, 113, 11–22 CrossRef.
  25. E. Suzuki, T. Komuro, Y. Kanno, M. Okazaki and H. Tobita, Facile 1,2-Migration of a Methyl Group on a {Dimethoxy(methyl)silyl}tungsten Complex: Formation of a Base-Stabilized (Dimethoxysilylene)(methyl) Complex, Organometallics, 2010, 29, 5296–5300 CrossRef CAS.
  26. E. Suzuki, T. Komuro, M. Okazaki and H. Tobita, Synthesis and Characterization of DMAP-Stabilized Aryl(silylene) Complexes and (Arylsilyl)(DMAP) Complexes of Tungsten: Mechanistic Study on the Interconversion between These Complexes via 1,2-Aryl Migration, Organometallics, 2009, 28, 1791–1799 CrossRef CAS.
  27. (a) J. R. Rumble, CRC handbook of chemistry and physics, 104th ed., CRC Press, Boca Raton, 2023 Search PubMed; (b) C. T. Watson, S. Cai, N. V. Shokhirev and F. A. Walker, NMR and EPR Studies of Low-Spin Fe(III) Complexes of meso-Tetra-(2,6-Disubstituted Phenyl)Porphyrinates Complexed to Imidazoles and Pyridines of Widely Differing Basicities, Inorg. Chem., 2005, 44, 7468–7484 CrossRef CAS PubMed.
  28. M. Okazaki, E. Suzuki, N. Miyajima, H. Tobita and H. Ogino, Facile Isomerization of a Tungsten Silyl Complex to a Base-Stabilized Silylene Complex via 1,2-Migration of an Aryl Group, Organometallics, 2003, 22, 4633–4635 CrossRef CAS.
  29. W. S. Sheldrick, The Chemistry of Organic Silicon Compounds, John Wiley & Sons, New York, 1989, Chapter 3 Search PubMed.
  30. M. Hirotsu, T. Nunokawa and K. Ueno, Synthesis and Reactivity of a Donor-Free (Silyl)(silylene)molybdenum Complex: Novel Insertion Reaction of an Isocyanide into a Si-C Bond, Organometallics, 2006, 25, 1554–1556 CrossRef CAS.
  31. (a) C. Zhao, R. M. Parrish, M. D. Smith, P. J. Pellechia, C. D. Sherrill and K. D. Shimizu, Do Deuteriums Form Stronger CH–π Interactions?, J. Am. Chem. Soc., 2012, 134, 14306–14309 CrossRef CAS PubMed; (b) M. Nishio, CH/π hydrogen bonds in crystals, CrystEngComm, 2004, 6, 130–158 RSC; (c) M. Nishio, The CH/π hydrogen bond in chemistry. Conformation, supramolecules, optical resolution and interactions involving carbohydrates, Phys. Chem. Chem. Phys., 2011, 13, 13873–13900 RSC; (d) M. Nishio, Y. Umezawa, J. Fantini, M. S. Weiss and P. Chakrabarti, CH–π hydrogen bonds in biological macromolecules, Phys. Chem. Chem. Phys., 2014, 16, 12648–12683 RSC; (e) M. Nishio, M. Hirota and Y. Umezawa, The CH/π interaction evidence, nature, and consequence, Wiley-VCH, New York, 1998, Chapter 6 Search PubMed; (f) F. A. Perras, D. Marion, J. Boisbouvier, D. L. Bryce and M. J. Plevin, Observation of CH⋯π Interactions between Methyl and Carbonyl Groups in Proteins, Angew. Chem., Int. Ed., 2017, 56, 7564–7567 CrossRef CAS PubMed.
  32. (a) S. Shimomura, M. Higuchi, R. Matsuda, K. Yoneda, Y. Hijikata, Y. Kubota, Y. Mita, J. Kim, M. Takata and S. Kitagawa, Selective sorption of oxygen and nitric oxide by an electron-donating flexible porous coordination polymer, Nat. Chem., 2010, 2, 633–637 CrossRef CAS PubMed; (b) S. Kusaka, Y. Itoh, A. Hori, J. Usuba, J. Pirillo, Y. Hijikata, Y. Ma and R. Matsuda, Adsorptive-dissolution of O2 into the potential nanospace of a densely fluorinated metal-organic framework, Nat. Commun., 2024, 15, 10117 CrossRef CAS PubMed.
  33. (a) H. P. G. Thompson and G. M. Day, Which conformations make stable crystal structures? Mapping crystalline molecular geometries to the conformational energy landscape, Chem. Sci., 2014, 5, 3173–3182 RSC; (b) S. L. Price, Control and prediction of the organic solid state: a challenge to theory and experiment, Proc. R. Soc. Ser. A, 2018, 474, 20180351 CrossRef PubMed; (c) S. E. Wright, J. C. Cole and A. J. Cruz-Cabeza, Conformational Change in Molecular Crystals: Impact of Solvate Formation and Importance of Conformational Free Energies, Cryst. Growth Des., 2021, 21, 6924–6936 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available: Synthetic and experimental procedures, characterisation data, details of X-ray crystallographic analysis, powder X-ray diffraction, micro-ED, thermogravimetric analysis, Le Bail analysis, gas sorption measurement and NMR spectra. CCDC 2445984 (4). For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d5qi01162k

This journal is © the Partner Organisations 2025
Click here to see how this site uses Cookies. View our privacy policy here.