DOI:
10.1039/D5SC04311E
(Edge Article)
Chem. Sci., 2025, Advance Article
Wrapping a single crystal spin-crossover complex with a single crystal non-spin-crossover complex to modulate the spin-transition temperature
Received
12th June 2025
, Accepted 24th October 2025
First published on 27th October 2025
Abstract
The formation of a heterojunction interface between different materials could lead to emergent functions that are not the simple sum of the properties of each component. Here we report the modulation of physical properties due to the lattice mismatch between two molecular crystals. Using metal complexes as a motif, we created a crystalline object that can be referred to as a core–shell crystal, in which a single crystal of one complex (FeII complex) is wrapped with a single crystal of a different complex (CoII or ZnII complex). X-ray analysis revealed that the FeII complex constituting the inner core exhibits spin-crossover behavior, while the CoII (or ZnII) complex constituting the outer shell does not. The crystal structure of the outer shell is different from that formed through spontaneous crystallization of the CoII (or ZnII) complex alone, but similar to that of the high-temperature phase of the spin-crossover FeII complex. Interestingly, the spin transition behavior of the FeII complex inside the core–shell crystals changes, demonstrating that in molecular materials, the formation of a heterojunction interface can modulate the properties of the entire bulk crystal. The fabrication of a ternary core–shell crystal using FeII, CoII and ZnII complexes is also presented.
Introduction
It is well known that in epitaxially grown films of inorganic materials, the lattice mismatch between two compounds leads to modulation of the physical properties of the material deposited on the substrate.1–3 Does this hold true for molecular crystals? This question appears to be non-trivial, since unlike inorganic materials in which atoms are infinitely connected by strong chemical bonds, these materials consist of discrete molecules assembled together by weak intermolecular forces. It is interesting to consider whether, for example, if two different molecular crystals were brought into close contact, the structural information of each component could influence the other, causing a change in the overall properties. In reality, however, it is not easy to create a material in which two types of molecular crystals are closely linked together at the molecular level. This is because the lattice mismatch between dissimilar molecular crystals is too large due to the complexity and anisotropy of their components. Hence, when growing crystals from a solution of a mixture of two compounds, fractional recrystallization usually occurs.
In the past, there have been several reports on the construction of heterostructures using metal–organic frameworks (MOFs),4–7 π-electronic molecules8–10 and metal complexes11–13 as constituents. However, little attention has been paid to the effect of lattice mismatches at heterojunction interfaces on the physical properties. In the present work, to create an elaborate heterojunction interface between two molecular crystals, we used metal complexes, which allow control over the physical properties without significantly changing the molecular structure by fixing the ligand while changing the type of metal ion. To explore the changes in physical properties due to the formation of heterojunction interfaces, we focused on the spin-crossover phenomenon, in which the spin state of a metal complex changes in response to external stimuli such as heat and light.14–17 Previous studies on nanoscale core–shell spin-crossover systems have reported that interfacial lattice strain can modulate the spin-transition temperature.18–24 However, this phenomenon has yet to be experimentally demonstrated in macroscopic single crystals. In the present work, we address this gap by demonstrating that even a small fraction of a heterojunction interface within a bulk crystal can modulate its overall spin-crossover behavior.
The above material design concept was embodied using metal complexes ligated with 4-carboxy-2,6-di(pyrazol-1-yl)pyridine25,26 (cdpp) with compositions of [M(cdpp)2](BF4)2 [Mcdpp (M = Fe, Co, Zn), Fig. 1]. Fecdpp undergoes a first-order phase transition between a low-spin state [total spin quantum number (S) = 0] and a high-spin state (S = 2) as the low-temperature and high-temperature phases, respectively, accompanied by a change in the molecular structure.25,26 On the other hand, Cocdpp has only a high-spin state (S = 3/2), and Zncdpp is diamagnetic in nature (S = 0), with both of them exhibiting no first-order phase transitions. We created a crystalline object that can be referred to as a core–shell crystal, in which a single crystal of Fecdpp is wrapped with a single crystal of Cocdpp or Zncdpp. A key to the successful formation of such core–shell crystals (Fecdpp@Cocdpp and Fecdpp@Zncdpp, Fig. 1b) lies in the seeded crystal growth method, which has been used to construct block supramolecular assemblies27–32 and block crystallizable polymer-based micelles.33–36 Interestingly, in this seeded crystallization, the crystal structure of Cocdpp or Zncdpp that forms the outer shell is different from that when it crystallizes alone. What is even more interesting is that the spin transition temperature of the Fecdpp core changes, where the greater the lattice mismatch between the core and shell crystals, the lower the spin transition temperature. Here we report the preparation, crystal structures and spin-crossover behavior of these core–shell crystals. The fabrication and characterization of a ternary core–shell crystal (Fecdpp@Cocdpp@Zncdpp, Fig. 1b) is also described.
 |
| | Fig. 1 (a) Synthetic scheme of Mcdpp (M = Fe, Co, Zn). cdpp: 4-carboxy-2,6-di(pyrazol-1-yl)pyridine. (b) Schematic illustration of core–shell crystals of Mcdpp using a seeded crystal growth method. | |
Results and discussion
Preparation and characterization of single crystals of Mcdpp
The 4-carboxy-2,6-di(pyrazol-1-yl)pyridine (cdpp) ligand was synthesized according to previously reported procedures.37 Single crystals of Mcdpp (M = Fe, Co, Zn) were prepared using a liquid–liquid diffusion crystallization method (Fig. S1, SI). Typically, an acetonitrile solution of a mixture of cdpp (1.0 eq.) and Fe(BF4)2·6H2O (0.5 eq.) is layered on top of a highly concentrated acetonitrile solution (6.6 M) of tetrabutylammonium tetrafluoroborate (TBABF4) and allowed to stand at 25 °C for two days, leading to the formation of single crystals of Fecdpp (Fig. 2a). By following a similar procedure as for Fecdpp, except that Co(BF4)2·6H2O (0.5 eq.) or Zn(BF4)2·nH2O (0.5 eq.) are used instead of Fe(BF4)2·6H2O, single crystals of Cocdpp or Zncdpp can be prepared (Fig. 2b and c).
 |
| | Fig. 2 POM images and X-ray crystal structures (molecular and packing structures) of (a) Fecdpp, (b) Cocdpp and (c) Zncdpp. Color code: carbon = gray, nitrogen = blue, oxygen = red, iron = orange, cobalt = purple, zinc = sky blue, fluorine = green, boron = pink. Hydrogen atoms have been omitted for clarity. | |
Fig. 2 shows the X-ray crystal structures of Fecdpp, Cocdpp and Zncdpp at 93 K, along with polarized optical micrographs (POM) of single crystal samples for each complex. The single crystal of Fecdpp belongs to a monoclinic system with space group C2/c (Fig. 2a). Fecdpp has an average Fe–N bond length (dFe–N) of 1.946 Å and an octahedral distortion parameter (Σ) of 85.32°. These values are typical for low-spin state FeII complexes.14–16,25,26 In terms of packing structure, the Fecdpp molecules assemble to form a one-dimensional (1D) hydrogen-bonded chain structure along the b-axis. When single crystalline Fecdpp is heated, it undergoes a phase transition involving a spin crossover with a change from a low- to a high-spin state. Thus, as determined from the X-ray crystal structure, the values of dFe–N (1.946 Å) and Σ (85.32°) at 93 K change at 393 K to 2.156 Å and 163.1°, respectively (Fig. S2, SI).
Cocdpp crystallizes into an orthorhombic system with space group Pbcn, in which a 1D hydrogen-bonded chain structure of Cocdpp is formed along the a-axis (Fig. 2b). In the crystal, the average Co–N bond length (dCo–N) of Cocdpp is 2.122 Å, and its Σ value is 135.4°, indicating that Cocdpp is in a high-spin state.38 Likewise, Zncdpp crystallizes into an orthorhombic space group (Pbcn), and the packing structure of the Zncdpp molecules is almost identical to that observed for single-crystalline Cocdpp (Fig. 2c).
Preparation and characterization of core–shell crystals
Using single crystals of the spin-crossover Fecdpp complex as seeds, we examined crystal growth of the Cocdpp complex from the seed surface (Fig. S1, SI). Thus, single crystals of Fecdpp [number-average area (An) = 8.07 × 103 μm2 and dispersity index (Aw/An) = 1.04] were immersed in an acetonitrile solution (6.6 M) of TBABF4. On top of this, a saturated acetonitrile solution of Cocdpp was layered slowly. After two days, core–shell crystals (denoted hereafter as Fecdpp@Cocdpp) were formed (Fig. 3a). An optical micrograph of the resulting crystals illustrates that the Fecdpp seed crystals are covered with an outer shell of crystalline Cocdpp (Fig. S3, SI), indicating quantitative formation of Fecdpp@Cocdpp (An = 1.21 × 104 μm2 and Aw/An = 1.02). Scanning electron microscopy energy dispersive X-ray spectroscopy (SEM-EDX) of Fecdpp@Cocdpp confirmed that the Fe content in the outer shell is below the detection limit (Fig. 3b). For analysis of the inner core, a Fecdpp@Cocdpp crystal was embedded in an epoxy resin and sliced using an ultramicrotome. The SEM-EDX profile of a sliced sample of Fecdpp@Cocdpp shows the presence of a distinct heterojunction interface between the Fecdpp core and the Cocdpp shell (Fig. 3c). These results demonstrate the formation of core–shell crystals having well-defined heterojunction interfaces.
 |
| | Fig. 3 (a) POM image of Fecdpp@Cocdpp. (b and c) SEM-EDX elemental maps (top) and total area traces (bottom) of Fecdpp@Cocdpp (b) before and (c) after slicing by a ultramicrotome. (d) Crystal structure of Cocdpp in the outer shell along the a-axis. (e) POM image of Fecdpp@Zncdpp. (f and g) SEM-EDX elemental maps (top) and total area traces (bottom) of Fecdpp@Zncdpp (f) before and (g) after slicing by a ultramicrotome. (h) Crystal structure of Zncdpp in the outer shell along the a-axis. | |
When the seeded crystallization event was monitored in real time, the formation of Fecdpp@Cocdpp was observed after leaving the two-layer solution to stand for about 5 h (Fig. S4, SI), which is faster than the spontaneous crystallization of Cocdpp (about 20 h) without seeds under otherwise identical conditions (Fig. S5, SI). Thus, the surface of the seed Fecdpp crystal may provide kinetic perturbation in the crystallization of Cocdpp.
The inner core and outer shell of the crystals were cut out and each segment analyzed by single-crystal X-ray analysis at 93 K. As expected, before and after the seeded crystal growth process, the crystal structure of the core Fecdpp remains unchanged. On the other hand, the crystal structure of the Cocdpp shell is different from when Cocdpp crystallizes alone in acetonitrile (Fig. 2b), and instead, it inherits the structural feature of the core Fecdpp (Fig. 3d and S6, SI). Thus, the Cocdpp forming the outer shell has the space group C2/c, where the difference in lattice constant with the core Fecdpp is within a range of 2.5% or less (Table S1, SI). This suggests that on the surface of the Fecdpp seed crystal, a kinetic assembly of Cocdpp is induced, which grows to become the outer shell. Notably, optical microscopy (OM) and scanning electron microscopy (SEM) have not shown any seams on the surfaces or edges of the core–shell crystals (Fig. S3 and S7, SI). In POM (Fig. S8, SI), the molecular orientations in the core and shell of Fecdpp@Cocdpp are identical to each other. These observations indicate that Fecdpp@Cocdpp is formed through a topotactic crystal growth process,39,40 resulting in a core–shell type “pseudo-single crystal” with two distinct well-defined domains.
Despite the difference in crystal structure and space group, there is no significant difference in the molecular structure of Cocdpp in the crystal formed in the absence or presence of the Fecdpp seed. For example, the values of dCo–N (2.118 Å) and Σ (134.1°) are virtually identical to those observed in the Cocdpp crystal alone (dCo–N = 2.122 Å and Σ = 135.4°), indicating that the Cocdpp shell crystal is in the high-spin state.38
Similar to Fecdpp@Cocdpp, using single crystals of Fecdpp (An = 3.63 × 103 μm2) as seeds, crystals of Zncdpp can grow around them (Fig. 3e). The resulting core–shell crystals, Fecdpp@Zncdpp (An = 6.05 × 103 μm2 and Aw/An = 1.06), have been unambiguously characterized by SEM-EDX (Fig. 3f and g) and single-crystal X-ray analysis (Fig. 3h and S9, SI). The Zncdpp shell of Fecdpp@Zncdpp adopts the C2/c space group (Fig. 3h), while crystals with the Pbcn space group grow from an acetonitrile solution of Zncdpp alone (Fig. 2c). The difference in lattice constant with the core Fecdpp is within a range of 3.1% or less (Table S1, SI). As in the case of Fecdpp@Cocdpp, the Zncdpp shell inherits the structural feature of the core Fecdpp (Fig. 3h and S9, SI).
Unfortunately, attempts to crystallize the Fecdpp complex from the seed surface of either Cocdpp or Zncdpp have not been successful. Presumably, Fecdpp would be unable to give Pbcn space group crystals as a kinetic form on the Cocdpp and Zncdpp crystal surfaces.
Spin-crossover behavior of core–shell crystals
Since spin-crossover is a first-order phase transition involving structural changes,14–16 its occurrence can be detected using differential scanning calorimetry (DSC) measurements. Prior to DSC measurements, we confirmed by thermogravimetric analysis (TGA) that all samples are thermally stable up to approximately 450 K, with no significant weight loss within the experimental temperature range (Fig. S10, SI). Fig. 4a shows the DSC profiles of Fecdpp, Fecdpp@Cocdpp and Fecdpp@Zncdpp. Upon heating from 223 K at a rate of 1 K min−1, Fecdpp exhibits an endothermic peak at 348 K due to a spin-crossover phenomenon. At the same temperature, the colour of the crystals changes from red to orange (Fig. S11, SI). Similarly, Fecdpp@Cocdpp and Fecdpp@Zncdpp undergo spin-crossover, with the appearance of endothermic peaks at 347 K and 346 K, respectively. While the difference in phase transition temperature is rather small, a clear difference can be observed between Fecdpp and the core–shell crystals. OM visualized the colour change occurring only in the cores of Fecdpp@Cocdpp and Fecdpp@Zncdpp (Fig. S11, SI).41,42 For comparison, we also measured DSC of a sample in which Fecdpp single crystals are wrapped in an epoxy resin and confirmed that there is no change in the phase transition temperature associated with the spin crossover of Fecdpp (Fig. S12, SI).
 |
| | Fig. 4 (a) DSC profiles of Fecdpp (black), Fecdpp@Cocdpp (red) and Fecdpp@Zncdpp (orange) upon heating at a rate of 1 K min−1. (b) Temperature-dependent magnetic susceptibility of Fecdpp (black), Fecdpp@Cocdpp (red) and Fecdpp@Zncdpp (orange) upon heating at a rate of 1 K min−1 under a magnetic field of 1 T (see Fig. S13 for the cooling process, SI). The χmT values were calculated assuming a 50 : 50 molar ratio between Fecdpp and Mcdpp (M = Co, Zn) in the core–shell crystals. Plots of spin-transition temperatures (T1/2) versus core/shell lattice constant mismatch at (c) b and (d) β. | |
Using a superconducting quantum interference device (SQUID) magnetometer, we investigated the temperature-dependence of the magnetic susceptibility of Fecdpp, Fecdpp@Cocdpp and Fecdpp@Zncdpp, upon heating from 4 K at a rate of 1 K min−1 under an external magnetic field of 1 T (Fig. 4b). Each SQUID measurement was repeated twice independently, which confirmed the reproducibility of the data. Consistent with previous reports,25,26 we observed that the χmT value of Fecdpp abruptly increases to 3.0 emu K mol−1 with a spin-transition temperature (T1/2) of 351 K, which is due to spin-crossover from the low- (S = 0) to high-spin state (S = 2). For Fecdpp@Cocdpp, the χmT value below 345 K was determined to be approximately 1.6 emu K mol−1, reflecting both high-spin state of Cocdpp (S = 3/2) and low-spin state of Fecdpp (S = 0). The χmT value greatly increases to 3.0 emu K mol−1 due to the spin-crossover behavior of the core Fecdpp with T1/2 of 349 K. Fecdpp@Zncdpp also displays spin-crossover behavior with T1/2 of 348 K (Fig. 4b). The fact that the T1/2 values for Fecdpp@Cocdpp (349 K) and Fecdpp@Zncdpp (348 K) are similar but apparently lower than that for Fecdpp (T1/2 = 351 K) shows that the wrapping of crystalline Fecdpp with other complex crystals perturbs the original spin-crossover behavior. It has been reported that the high-spin state of FeII complexes is stabilized in solid-solution systems diluted with NiII, CoII, or ZnII ions, which have ionic radii similar to that of FeII in the high-spin state.43–45 However, it is notable that the phase-transition behavior of the entire bulk crystal changes when spatially distinct crystalline domains are closely linked together.
We found that the degree of lattice constant mismatch of each block segment is related to T1/2. In Fecdpp@Cocdpp, when comparing the values of b and β between the Fecdpp core and the Cocdpp shell, there are differences of 2.42% and 1.91%, respectively (Fig. S14 and Table S1, SI). Fecdpp@Zncdpp has differences in b and β of 3.02% and 2.02%, respectively (Table S1, SI), which are slightly larger than the case of Fecdpp@Cocdpp. Clearly, the greater the lattice mismatch between the core and shell crystals, the lower the spin transition temperature (Fig. 4c, d and S15, SI). It should be noted that the values of b and β of Cocdpp and Zncdpp, which form the outer shell, are closer to those of the high temperature phase of Fecdpp (i.e., high-spin state), than to those of its low temperature phase (i.e., low-spin state).
Since the interface of the core–shell crystals would be in a higher energy state than the interior of the Fecdpp core, the phase transition is thought to be initiated from the interface. The interface is covered by the crystalline shell of Cocdpp or Zncdpp with a structure similar to the high-temperature phase of Fecdpp. That is, for the low-temperature phase of Fecdpp before the transition, the shell has a structural mismatch, while for the high-temperature phase of Fecdpp generating after the transition, the shell has a structural similarity. This feature of the heterojunction interface is considered responsible for lowering the phase transition temperature in the core–shell crystal systems. In fact, the structure of the shell Zncdpp is more similar to the high-temperature phase of Fecdpp than that of the Cocdpp shell, and Fecdpp@Zncdpp undergoes the phase transition at a lower temperature than Fecdpp@Cocdpp. Such interfacial effects have also been discussed previously in nanoscale coordination polymer core–shell systems, where the modulation of spin-transition behavior is understood in terms of elastic effects due to a mismatch in lattice parameters at the interface.18–24 The present results may extend this nanoscale understanding to macroscopic core–shell structures composed of molecular crystals.
Preparation and characterization of a ternary core–shell crystal
We have shown that sequential seeded crystallization allows for the preparation of an intriguing ternary core–shell crystals. As mentioned above, crystalline Cocdpp and Zncdpp can be grown from the surface of a seed crystal of Fecdpp, and also both can form a crystal structure with the monoclinic C2/c space group. Taking advantage of these structuring properties, Fecdpp@Cocdpp@Zncdpp can be obtained through the sequential seeded crystallization of Cocdpp and Zncdpp from a single crystal of Fecdpp as seed (Fig. 1b). The ternary core–shell crystals have been characterized by OM and SEM-EDX (Fig. 5a, b and S16, SI). In DSC, Fecdpp@Cocdpp@Zncdpp, upon heating, exhibits an endothermic peak at 347 K (Fig. S12, SI), which is the same as in the case of Fecdpp@Cocdpp. As expected, the phase transition behavior of the core Fecdpp is affected only by the outer shells in direct contact with it.
 |
| | Fig. 5 (a) POM image of Fecdpp@Cocdpp@Zncdpp. (b) SEM-EDX elemental map and total area trace along one axis of Fecdpp@Cocdpp@Zncdpp after slicing by a ultramicrotome. Color code: iron = red, cobalt = green, zinc = sky blue. | |
Conclusions
We have presented the preparation and characterization of core–shell crystals, in which a single crystal spin-crossover complex (Fecdpp) is wrapped with a single crystal of a non-spin-crossover metal complex (Cocdpp or Zncdpp), using the seeded crystallization approach. Interestingly, the crystal structures of the Cocdpp and Zncdpp shells are different from those formed by their spontaneous crystallization, and instead they inherit the crystal structure of the inner Fecdpp core. When Fecdpp is surrounded by an outer shell of Cocdpp or Zncdpp, its phase transition shifts to a lower temperature, with the degree of variation becoming larger the greater the similarity between the lattice constants of Cocdpp or Zncdpp, and the structure of the high temperature phase of Fecdpp. In this seeded crystallization, the Fecdpp seed induces the outer shell complex to crystallize into a structure different from its intrinsic one, and thus indirectly tunes the inherent phase transition behavior. The present study, which demonstrates that a heterojunction interface in different molecular crystals can have a noticeable effect on the physical properties of the bulk crystals, sparks new interest in the seeded crystallization approach to interface design of molecular crystals.
Author contributions
To. F. conceived the project; To. F. and Ta. F. designed the experiments; M. T., N. M. and To. F. carried out the experiments and analysed the data; To. F. and Ta. F. co-wrote the manuscript.
Conflicts of interest
There are no conflicts to declare.
Data availability
The data supporting this article have been included as part of the supplementary information (SI). The supplementary information includes crystal data, OM/POM images, TGA data, DSC profiles, SEM images and magnetic susceptibility data. See DOI: https://doi.org/10.1039/d5sc04311e.
CCDC 2214914 (for Fecdpp), 2408263 (for Fecdpp at 393 K), 2214916 (for Cocdpp), 2214918 (for Zncdpp), 2214915 (for Cocdpp in the shell) and 2214919 (for Zncdpp in the shell) contain the supplementary crystallographic data for this paper.46a–f
Acknowledgements
This work was supported by Japan Science and Technology Agency (JST) ACT-X (JPMJAX24DH for To.F.), JSPS KAKENHI (JP23K13728 for To.F.) and Grant-in-Aid for Transformative Research Areas (A) “Supra-ceramics” (JP23H04617 for To.F.). This work was also supported by “Crossover Alliance to Create the Future with People, Intelligence and Materials” from MEXT, Japan. We thank the Materials Analysis Division, Core Facility Center, Institute of Science Tokyo, for their support with SEM-EDX and single-crystal X-ray diffraction measurements.
Notes and references
- S. A. Chambers, Surf. Sci. Rep., 2000, 39, 105–180 CrossRef CAS.
- J. Mannhart and D. G. Schlom, Science, 2010, 327, 1607–1611 CrossRef CAS PubMed.
- H. Y. Hwang, Y. Iwasa, M. Kawasaki, B. Keimer, N. Nagaosa and Y. Tokura, Nat. Mater., 2012, 11, 103–113 CrossRef CAS PubMed.
- S. Furukawa, K. Hirai, K. Nakagawa, Y. Takashima, R. Matsuda, T. Tsuruoka, M. Kondo, R. Haruki, D. Tanaka, H. Sakamoto, S. Shimomura, O. Sakata and S. Kitagawa, Angew. Chem., Int. Ed., 2009, 48, 1766–1770 CrossRef CAS PubMed.
- M. Pan, Y.-X. Zhu, K. Wu, L. Chen, Y.-J. Hou, S.-Y. Yin, H.-P. Wang, Y.-N. Fan and C.-Y. Su, Angew. Chem., Int. Ed., 2017, 56, 14582–14586 CrossRef CAS PubMed.
- C. Liu, Q. Sun, L. Lin, J. Wang, C. Zhang, C. Xia, T. Bao, J. Wan, R. Huang, J. Zou and C. Yu, Nat. Commun., 2020, 11, 4971 CrossRef CAS PubMed.
- K. Leng, H. Sato, Z. Chen, W. Yuan and T. Aida, J. Am. Chem. Soc., 2023, 145, 23416–23421 CrossRef CAS PubMed.
- M.-P. Zhuo, J.-J. Wu, X.-D. Wang, Y.-C. Tao, Y. Yuan and L.-S. Liao, Nat. Commun., 2019, 10, 3839 CrossRef PubMed.
- Y. Su, B. Wu, S. Chen, J. Sun, Y. Yu, M. Zhuo, Z. Wang and X. Wang, Angew. Chem., Int. Ed., 2022, 61, e202117857 CrossRef CAS PubMed.
- Q. Lv, X.-D. Wang, Y. Yu, M.-P. Zhuo, M. Zheng and L.-S. Liao, Nat. Commun., 2022, 13, 3099 CrossRef CAS PubMed.
- C. R. R. Adolf, S. Ferlay, N. Kyritsakas and M. W. Hosseini, J. Am. Chem. Soc., 2015, 137, 15390–15393 CrossRef CAS PubMed.
- M. Wakizaka, S. Kumagai, H. Wu, T. Sonobe, H. Iguchi, T. Yoshida, M. Yamashita and S. Takaishi, Nat. Commun., 2022, 13, 1188 CrossRef CAS PubMed.
- Q. Wan, M. Wakizaka, N. Funakoshi, Y. Shen, C.-M. Che and M. Yamashita, J. Am. Chem. Soc., 2023, 145, 14288–14297 CrossRef CAS PubMed.
- M. A. Halcrow, Chem. Soc. Rev., 2011, 40, 4119–4142 RSC.
- P. Guionneau, Dalton Trans., 2013, 43, 382–393 RSC.
- G. Molnár, S. Rat, L. Salmon, W. Nicolazzi and A. Bousseksou, Adv. Mater., 2018, 30, 1703862 CrossRef PubMed.
- M. Gavara-Edo, R. Córdoba, F. J. Valverde-Muñoz, J. Herrero-Martín, J. A. Real and E. Coronado, Adv. Mater., 2022, 34, e2202551 CrossRef PubMed.
- L. Catala, D. Brinzei, Y. Prado, A. Gloter, O. Stéphan, G. Rogez and T. Mallah, Angew. Chem., Int. Ed., 2008, 48, 183–187 CrossRef PubMed.
- C. M. Quintero, G. Félix, I. Suleimanov, J. S. Costa, G. Molnár, L. Salmon, W. Nicolazzi and A. Bousseksou, Beilstein J. Nanotechnol., 2014, 5, 2230–2239 CrossRef PubMed.
- Y.-X. Wang, D. Qiu, S.-F. Xi, Z.-D. Ding, Z. Li, Y. Li, X. Ren and Z.-G. Gu, Chem. Commun., 2016, 52, 8034–8037 RSC.
- H. Oubouchou, Y. Singh and K. Boukheddaden, Phys. Rev. B, 2018, 98, 014106 CrossRef.
- C. Felts, A. Slimani, J. M. Cain, M. J. Andrus, A. R. Ahir, K. A. Abboud, M. W. Meisel, K. Boukheddaden and D. R. Talham, J. Am. Chem. Soc., 2018, 140, 5814–5824 CrossRef PubMed.
- K. Affes, A. Slimani, A. Maalej and K. Boukheddaden, Chem. Phys. Lett., 2019, 718, 46–53 CrossRef CAS.
- N. Natt and B. Powell, 2019, ChemRxiv, preprint, DOI:10.26434/chemrxiv-2025-tgqb3.
- A. Abhervé, M. Clemente-León, E. Coronado, C. J. Gómez-García and M. López-Jordà, Dalton Trans., 2014, 43, 9406–9409 RSC.
- V. García-López, M. Palacios-Corella, A. Abhervé, I. Pellicer-Carreño, C. Desplanches, M. Clemente-León and E. Coronado, Dalton Trans., 2018, 47, 16958–16968 RSC.
- B. Adelizzi, N. J. V. Zee, L. N. J. de Windt, A. R. A. Palmans and E. W. Meijer, J. Am. Chem. Soc., 2019, 141, 6110–6121 CrossRef CAS PubMed.
- W. Zhang, W. Jin, T. Fukushima, A. Saeki, S. Seki and T. Aida, Science, 2011, 334, 340–343 CrossRef CAS PubMed.
- S. H. Jung, D. Bochicchio, G. M. Pavan, M. Takeuchi and K. Sugiyasu, J. Am. Chem. Soc., 2018, 140, 10570–10577 CrossRef CAS PubMed.
- W. Wagner, M. Wehner, V. Stepanenko and F. Würthner, J. Am. Chem. Soc., 2019, 141, 12044–12054 CrossRef CAS PubMed.
- Q. Wan, W.-P. To, X. Chang and C.-M. Che, Chem, 2020, 6, 945–967 CAS.
- N. Sasaki, J. Kikkawa, Y. Ishii, T. Uchihashi, H. Imamura, M. Takeuchi and K. Sugiyasu, Nat. Chem., 2023, 15, 922–929 CrossRef CAS PubMed.
- L. R. MacFarlane, H. Shaikh, J. D. Garcia-Hernandez, M. Vespa, T. Fukui and I. Manners, Nat. Rev. Mater., 2021, 6, 7–26 CrossRef CAS.
- Z. M. Hudson, D. J. Lunn, M. A. Winnik and I. Manners, Nat. Commun., 2014, 5, 3372 CrossRef PubMed.
- H. Qiu, Z. M. Hudson, M. A. Winnik and I. Manners, Science, 2015, 347, 1329–1332 CrossRef CAS PubMed.
- X.-H. Jin, M. B. Price, J. R. Finnegan, C. E. Boott, J. M. Richter, A. Rao, S. M. Menke, R. H. Friend, G. R. Whittell and I. Manners, Science, 2018, 360, 897–900 CrossRef CAS PubMed.
- T. Vermonden, D. Branowska, A. T. M. Marcelis and E. J. R. Sudhölter, Tetrahedron, 2003, 59, 5039–5045 CrossRef CAS.
- M. Nakaya, W. Kosaka, H. Miyasaka, Y. Komatsumaru, S. Kawaguchi, K. Sugimoto, Y. Zhang, M. Nakamura, L. F. Lindoy and S. Hayami, Angew. Chem., Int. Ed., 2020, 59, 10658–10665 CrossRef CAS PubMed.
- S. Kittaka, H. Hamaguchi, T. Umezu, T. Endoh and T. Takenaka, Langmuir, 1997, 13, 1352–1358 CrossRef CAS.
- P. Zhang, T. Ochi, M. Fujitsuka, Y. Kobori, T. Majima and T. Tachikawa, Angew. Chem., Int. Ed., 2017, 56, 5299–5303 Search PubMed.
- F. Varret, A. Slimani, K. Boukheddaden, C. Chong, H. Mishra, E. Collet, J. Haasnoot and S. Pillet, New J. Chem., 2011, 35, 2333–2340 RSC.
- M. Sy, D. Garrot, A. Slimani, M. Páez-Espejo, F. Varret and K. Boukheddaden, Angew. Chem., Int. Ed., 2016, 55, 1755–1759 CrossRef CAS PubMed.
- J.-P. Martin, J. Zarembowitch, A. Dworkin, J. G. Haasnoot and E. Codjovi, Inorg. Chem., 1994, 33, 2617–2623 CrossRef CAS.
- M. A. Halcrow, H. B. Vasili, C. M. Pask, A. N. Kulak and O. Cespedes, Dalton Trans., 2024, 53, 6983–6992 RSC.
- M. A. Halcrow, Dalton Trans., 2024, 53, 13694–13708 RSC.
-
(a) CCDC 2214914: Experimental Crystal Structure Determination, 2025, DOI:10.5517/ccdc.csd.cc2dbsvd;
(b) CCDC 2214915: Experimental Crystal Structure Determination, 2025, DOI:10.5517/ccdc.csd.cc2dbswf;
(c) CCDC 2214916: Experimental Crystal Structure Determination, 2025, DOI:10.5517/ccdc.csd.cc2dbsxg;
(d) CCDC 2214918: Experimental Crystal Structure Determination, 2025, DOI:10.5517/ccdc.csd.cc2dbszj;
(e) CCDC 2214919: Experimental Crystal Structure Determination, 2025, DOI:10.5517/ccdc.csd.cc2dbt0l;
(f) CCDC 2408263: Experimental Crystal Structure Determination, 2025, DOI:10.5517/ccdc.csd.cc2ltzxb.
|
| This journal is © The Royal Society of Chemistry 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.