Open Access Article
Yukina Suzuki
*a,
Nina Strasser
b,
Ryota Sakamoto
a and
Masahiro Yamashita
*acd
aDepartment of Chemistry, Graduate School of Science, Tohoku University, 6-3 Aramaki-Aza-Aoba, Aoba-Ku, Sendai 980-8578, Japan. E-mail: suzuki.yukina.s2@dc.tohoku.ac.jp; masahiro.yamashita.c5@tohoku.ac.jp
bInstitute of Solid State Physics, NAWI Graz, Graz University of Technology, Petersgasse 16, 8010 Graz, Austria
cSchool of Chemical Science and Engineering, Tongji University, Siping Road 1239, Shanghai 200092, P. R. China
dInstitute of Materials Research, Tohoku University, 2-1-1 Katahira, Aoba-Ku, Sendai 980-8577, Japan
First published on 13th January 2026
Employing metal–organic frameworks (MOFs) has been shown to be an effective strategy for extending spin–lattice relaxation times (T1) by modifying the vibrational properties of molecular spin qubits. Although a previous study employed terahertz spectroscopy to probe vibrational properties, the specific types of vibrational modes that affect T1 have remained unclear. In this study, we use periodic density functional theory calculations to investigate the vibrational properties of the MOF [{M(TCPP)}Zn2(H2O)2] (M = Cu, VO, and TCPP = tetrakis(4-carboxyphenyl)porphyrin) and the corresponding molecular crystals MTPP (TPP = tetraphenylporphyrin). Although the longer T1 of the vanadyl MOF compared to the VOTPP crystal has been attributed to the absence of low-frequency modes, our combined experimental Raman spectra and phonon simulations show that the mere presence of low-frequency modes does not necessarily lead to faster relaxation times. To rationalize the differences in T1 between the MOFs and the corresponding molecular crystals, we calculated the spin–phonon coupling (SPC) for each Γ-point phonon mode. Furthermore, we analysed correlations with the magnitude of the SPCs, the symmetry of the modes, and the in-plane and out-of-plane distortion of the porphyrin framework. Our results reveal that no single descriptor (frequency, symmetry, or distortion magnitude) can reliably predict the SPC strength, highlighting the multifactorial nature of the SPCs in these systems. This complexity underscores the importance of explicit computational treatments for identifying the key phonon modes that drive spin–lattice relaxation, while spectroscopic techniques such as low-frequency vibrational spectroscopy can provide complementary validation and qualitative insights.
Spin–lattice relaxation occurs mainly via two mechanisms: phonon-mediated processes and processes arising from spin–spin interactions (cross relaxation). In a magnetically well-diluted system and at higher temperatures,8 the phonon-mediated process becomes the primary relaxation mechanism. This is because the interactions between paramagnetic centres become negligible, and the cross relaxation is temperature-independent.10 In the phonon-mediated mechanism, different processes dominate in different temperature ranges. At low temperatures (below 20 K),11 the one-phonon-mediated direct process is typically dominant. In this process, the energy of the spin is directly exchanged with the phonon energy, which requires a precise energy match of the spin and the phonon. In typical electron paramagnetic resonance (EPR) experiments performed at X or Q-band, the energy difference of the total spin quantum number S = 1/2 (with magnetic sublevels ms = ±1/2) is less than a few cm−1, therefore largely limiting the available phonons for energy exchange. As the temperature increases, the two-phonon Raman process becomes dominant. The Raman process does not require a direct match of the phonon energy and the spin energy; only the difference between two phonon modes and the spin energy must match. This significantly loosens the requirement for spin–lattice relaxation and allows phonons with frequencies larger than a few cm−1 to participate in the process. Combined with the increased population of higher-energy phonons with increasing temperature, the two-phonon processes become the dominant relaxation mechanism typically above 20 K.12 Another crucial relaxation mechanism is the local process. This process is also classified as a two-phonon process, similar to the Raman process,11 but involves optical phonons or vibrations with localized character and is typically identified as the major relaxation mechanism at higher temperatures.13,14
Given that these phonon-mediated processes, particularly the Raman and local mode processes, make dominant contributions to the spin–lattice relaxation, tuning the vibrational spectrum of molecular qubits can be an effective strategy for suppressing spin–lattice relaxation at elevated temperatures. Indeed, reducing low-frequency vibrations has been shown to be an effective strategy to mitigate spin–lattice relaxation. A common approach to suppress such low-frequency vibrations is to create a rigid ligand environment, thereby shifting the vibrational modes around the spin centre to higher energy.13,15 In addition to rigidifying the ligand environment, the vibrational spectrum can also be tuned by embedding molecular units into an extended framework. This approach, which is the focus of this study, allows control over lattice vibrations through the surrounding host structure. In this context, a previous study7 demonstrated that incorporating molecular units into the lattice of a metal–organic framework (MOF) can prolong the spin–lattice relaxation times, particularly at elevated temperatures. Terahertz (THz) spectroscopy suggested that the MOF lattice shifts optical phonons in the corresponding molecular crystal to higher energies, thereby reducing the efficiency of the spin–lattice relaxation. While the study by Yamabayashi et al.7 established the beneficial effect of framework incorporation on spin–lattice relaxation, the underlying vibrational contributions remained unexplored. Computational studies16,17 on spin–phonon coupling in vanadyl and copper molecular spin qubits in the gas phase suggest that only a few symmetry-allowed key vibrations have a significant impact on the spin–lattice relaxation time. Therefore, understanding how the MOF framework modifies these key vibrations in molecular solids is crucial for the rational design of MOFs with improved spin–lattice relaxation times.
In order to better understand the difference between the vibrational properties of qubits in MOFs and molecular crystals, we investigated prototypical MOFs with copper(II) and vanadyl(IV) porphyrins ([{Cu(TCPP)}Zn2(H2O)2] (TCPP = tetrakis(4-carboxyphenyl)porphyrin) (1) and [{VO(TCPP)}Zn2(H2O)2] (2)). These MOFs have previously been shown to have longer T1 than their corresponding molecular crystals of CuTPP (TPP = tetraphenylporphyrin) (3) and VOTPP (4)7,18–20 (Fig. S1) and serve therefore as an ideal model system for a comparative analysis (Fig. 1). The spin–lattice relaxation process is strongly influenced by both the electronic structure and the phonons. However, several studies7,18,19 have shown that the g factors are very similar between the MOF-based porphyrins and their molecular crystal analogues (Tables S1 and S2). The g factor describes how an electron's magnetic moment couples to an external magnetic field and is closely related to the electronic structure. The deviation of the g factor from the free-electron value (ge ∼ 2.0023) occurs due to spin–orbit coupling and the ligand environment. The magnitude of deviation of g factors from the free electron value is known to be a useful proxy for the size of the spin–orbit coupling.11 Thus, similar g factors suggest that phonons are the primary source of their distinct spin–lattice relaxation behaviours rather than the difference in their electronic structures.
![]() | ||
Fig. 1 The crystal structures of the copper(II) and vanadyl(IV) porphyrin systems studied in this work visualized using Mercury.22 View along the a axis of ([{Cu(TCPP)}Zn2(H2O)2]) (P2/m) (1) and [{VO(TCPP)}Zn2(H2O)2] (2). View along the c axis of CuTPP (TPP = tetraphenyl porphyrin) (I 2d) (3) and VOTPP (I4) (4). The unit cells are outlined with thin lines. Colour code: Cu – orange, V – silver, Zn – purple, N – light purple, O – red, C – grey, H – omitted for clarity. | ||
To probe this, the Γ-point phonons of systems (1)–(4) were calculated and compared using periodic density functional theory (DFT), as simplified single molecule or cluster models fail to reproduce low-frequency (terahertz) vibrations because they are highly sensitive to intermolecular and lattice interactions.21 Since the vibrational modes that dominate spin–lattice relaxation at relevant temperatures (room temperature or below) are low-frequency modes below 400 cm−1,16 where intermolecular interactions play a significant role, a precise description of these phonon modes is crucial. In addition to phonon calculations, spin–phonon coupling analysis was conducted in order to evaluate how these vibrational modes interact with the spin degrees of freedom and to understand the origin of their different spin-relaxation behaviours.
Raman spectra were collected at wavelengths of 488 nm and 532 nm with an NRS-5100 spectrometer to confirm that wavelength-dependent resonance effects did not influence the relative band intensities, enabling a direct comparison with simulated non-resonant spectra. At 532 nm, the measurable range was limited to >100 cm−1 due to the cutoff of the notch filter required to suppress Rayleigh scattering.
For (2), no single-crystal X-ray structure has been reported, as the structure determination was hampered by their twinning or polycrystalline nature.19 However, the previous work established that this system is assumed to be isostructural with (1).19 Therefore, its structural model was constructed by substituting Cu with an oxovanadium (VO) unit in (1). Because the VO unit breaks the twofold axis present in (1), the symmetry is reduced to P1, effectively lowering the lattice system from monoclinic to triclinic. To avoid artificial distortions of the lattice during structural optimization of (2), the lattice constants were kept fixed in this case, and only the atomic positions were relaxed.
For (4), the orientation of the VO group is disordered in the experimental single-crystal structure with the I4/m space group (CCDC: 112597634 or 22901937). Explicitly accounting for this disorder would require large supercells and is therefore not practical. To approximate the two limiting configurations, the unit cell was optimized in the P4/n (mirror plane lost, inversion preserved via a glide plane) and I4 (both inversion and mirror symmetry lost) space groups, as shown in as shown in Fig. S8. Geometry optimization and phonon calculation using the P4/n cell resulted in two imaginary frequency modes (∼−25 cm−1). Therefore, the structure with the I4 unit cell was employed for further discussion, and the symmetry analysis was only conducted based on this space group.
For the simulations with CRYSTAL23, the frequency calculations (phonon calculations at the Γ-point) were carried out based on the fully optimized structures (including lattice relaxations). Then, the computed frequencies were checked to ensure there are no imaginary modes to confirm that the optimized structure is stable. Subsequently, the intensities of the infrared (IR) and Raman modes were calculated in order to compare the spectra with our experimental measurements. Raman intensities were computed at the temperature of measurement (298 K) and the wavelength of the incident laser (488 nm and 532 nm). For simulations with VASP, phonon calculations were carried out using the finite displacement method implemented in Phonopy, employing the primitive unit cells.32,33 The amplitude of the atomic displacement was set to 0.01 Å. The convergence of the computed phonon modes was confirmed by the agreement of the eigenvectors (vide infra).
MOF are well known for their ability to adsorb solvent or gas molecules within their pores. In passing we note that in the simulations presented in this study, we used MOF models without any gas or solvent molecules in the pores. We present several justifications for this choice. First, the MOFs synthesized were solvent exchanged in acetone and then dried under vacuum. For EPR experiments to measure the relaxation time, these solid samples were sealed under high vacuum in a glass tube. This should effectively remove weakly bound molecules from the pores. Raman spectra were measured under ambient condition. In general, the phonons or vibrational modes in the THz region is very sensitive to any structural changes. Some MOFs have been reported to show noticeable difference in their vibrational spectra depending on the existence of absorbent molecules in the pore.52,53 The agreement between the experimental spectra and the simulated spectra in this study without considering any absorbents suggest that the model in the simulation is sufficiently accurate.
For the simulations using CRYSTAL23, various basis sets were tested ranging from double-ζ-quality basis sets (sol-def2-mSVP,55 6-31G**56–59) to triple-ζ basis sets (POB-TZVP,60 VTZP61). Among these, the composite method PBEsol0-3c/sol-def2-mSVP55 reproduced the spectral feature and lattice constants very well (Fig. 2, Fig. S10–S12 and Tables S5–S7). In passing, we note that PBEsol0-3c55 is a solid-state composite method based on the PBEsol0 hybrid functional,62 combined with a compact sol-def2-mSVP55 basis set and corrections for dispersion interactions on the D3 level63 using the Becke–Johnson damping function64 and basis-set superposition error. The method is known for its efficiency while maintaining good accuracy.55
However, subsequent eigenvector analysis (Fig. S13, SI) revealed that the basis sets of double-ζ-quality are not sufficient in reaching convergence, especially at low-frequency modes. In this analysis, the similarity of the eigenvectors (displacement vector of each atom) of the phonon modes computed with different functionals or basis sets were compared. The similarity was evaluated by taking the dot product of normalized displacement vectors for each atom.65 When two phonon modes from different calculations are equivalent, their normalized dot product approaches 1; conversely, values close to 0 indicate little or no similarity. Fig. S13b and d shows that the dot products with a well converged VASP calculation (PBE-D3) of the eigenvectors are less than 0.5 for various modes at low frequency, indicating that the phonon modes computed with double-ζ-quality basis sets (sol-def2-mSVP and 6-31G**) are not consistent. The convergence of VASP calculations were assumed from an excellent agreement of calculations with two different functionals (PBE-D3 and PBEsol-D3), as shown in Fig. S13a. Simulations with double-ζ-quality basis sets resulted in discrepancies in vibrational modes that show significant spin–phonon coupling coefficients in the subsequent spin–phonon coupling analysis (vide infra, Fig. S14 and S15). In order to gain consistent results across different methods, the VTZP basis set, which has previously been used for a Cu-based MOF and showed reasonable convergence,66 was required.
Upon increasing the basis set to this level, the agreement of the eigenvector overlaps across different calculations significantly improved. As shown in Fig. S13e and f, the PBE-D3/VTZP level calculations performed using CRYSTAL23 and VASP (PBE67 or PBEsol62 with ENCUT = 850 eV) gave nearly the same results for the phonon calculations and subsequent spin–phonon coupling calculations (vide infra, Fig. S17 and S18), validating the convergence of our phonon calculations. POB-TZVP60 is also a triple-ζ-quality basis set, it is known to give inconsistent results and tends to overestimate intermolecular forces, leading to contracted lattice parameters and blue-shifted vibrational modes.68 In line with this observation, the POB-TZVP basis set performed poorly here as well, producing the largest deviation in the lattice constants (Tables S6 and S7) and phonon eigenvector overlaps with those obtained from other methods (Fig. S13).
Within the spin-Hamiltonian framework, several approaches exist for computing SPC coefficients, depending on whether the g-tensor (g; Zeeman interaction) or the hyperfine tensor (A; hyperfine interaction) is considered to be the leading spin–lattice relaxation mechanism, and whether first or second derivatives with respect to normal coordinates (Q) are taken. Previously reported methods include the first derivative
(ref. 13, 16, 20, 38–41 and 69–72),
(ref. 41 and 73), the second derivative
(ref. 17), second mixed derivative
(ref. 11 and 73–75) and
(ref. 73 and 74). The choice of g or A to be differentiated depends on the assumed dominant mechanism of spin–lattice relaxation, which remains under debate.11,16 Despite this ongoing debate, a previous study has shown that g and A derivatives yield consistent trends in spin–phonon coupling behaviour, giving essentially the same vibrational modes as the dominant contributors.73
In this study, we employed the more widely adopted g-tensor approach to gain insights into the modes that have a substantial contribution to spin–lattice relaxation. Our primary objective is to rationalize the different relaxation behaviours of molecular crystals and MOFs bearing the same molecular units, rather than reproducing absolute T1 values. The g-tensor framework provides a particularly transparent connection between vibrations and magnetic relaxation, as g-tensor anisotropy arises directly related to spin–orbit coupling.16 We note that a complete and quantitative description of spin–lattice relaxation generally requires consideration of both g- and A-tensor modulation. However, because the g or A-tensor approach is reported to give the same trend,73 the choice between the g or A-tensor approach is unlikely to impact our qualitative final conclusions.
Fig. 3 presents the first-order SPC coefficients of the systems (1) and (3) determined by fitting g-values as a function of atomic displacement for each phonon mode. First-order spin–phonon coupling coefficients for the systems (2) and (4) are calculated similarly and shown in Fig. S27. Consistent with previous findings,16,40,70 only a small fraction of vibrational modes exhibits significant spin–phonon coupling, while the majority contribute negligibly. For (1), vibrational modes around 200 cm−1 and 380 cm−1 show the largest first-order SPC coefficients (Fig. 3) due to both correct symmetry and large displacement of the ligand (Table S13, see Discussion). In the case of (3), the dominant contributions arise from modes around 400 cm−1 similar to (1), with additional smaller contributions from around 120 cm−1, 200 cm−1, and 330 cm−1 modes (Fig. 3). For the vanadyl systems, both (2) and (4) have major first-order spin phonon coupling coefficients around 20–30 cm−1, 100 cm−1, 200 cm−1, 300 cm−1 and 400 cm−1 (Fig. S27). System (2) also exhibits additional contributions near 250 cm−1 and 350 cm−1. While small variations exist in the precise frequencies and number of contributing modes, the overall trend looks similar, with major contributions clustered around comparable frequency ranges (highlighted in green in Fig. 3).
The magnitudes of the SPC coefficients are also comparable between the MOF and molecular crystal counterparts. For (1), the largest coefficients are approximately 4 × 10−6 (gx and gy) and 5 × 10−5 (gz) (Fig. 3). For (3), the maximum values are about 1.5 × 10−6 (gx and gy) and 1.5 × 10−5 (gz). A similar trend is observed for the vanadyl systems of (2) and (4): the dominant SPC coefficients are on the order of 10−8 for gx and gy, and 10−6 for gz (Fig. S27). These similarities stand in contrast to the experimentally observed differences in spin–lattice relaxation between the MOF and molecular crystal systems. Typically, experimental T1 values associated with parallel components (gz) are larger, indicating a weaker spin–phonon coupling compared to that measured in the perpendicular (gx and gy) direction.81 This behaviour, however, contradicts the calculated trend, where gz (∼10−6) show larger SPC coefficients than gx and gy (∼10−8), for example in (2) and (4).
SPC coefficients set the intrinsic coupling strength, but phonon populations also control each mode's contribution to the spin–lattice relaxation at finite temperatures. To incorporate phonon population effects, SPC coefficients were scaled using the Bose–Einstein distribution (Fig. S29–S32). As shown in Fig. 4a, the theoretical temperature dependence of a value proportional to T1 derived from the summed contributions of all vibrational modes is nearly identical for (1) and (3), in apparent contrast to the experimental results (Fig. S1). For a quantitative comparison with the experiment, an empirical arbitrary factor is used to scale to the experimental value in the previous study.16 Since this method is a simplified method16 from the more complicated but rigorous method,80 we avoid a quantitative argument of T1 here but restrict our argument for only the qualitative trend. Nevertheless, the discrepancy between the computational and experimental results is clearly seen even at this qualitative level.
Fig. 5 presents the computed second-order (non-cross terms only) spin–phonon coupling coefficients. In contrast to the previous assumption that the first-order SPC shows a similar trend with the second-order SPC,16,69 our results reveal markedly different distributions of coupling strengths between the first- and second-order coefficients across all systems investigated (Fig. 3, 5 and Fig. S27, S28). Moreover, contrary to the first-order SPC, where the differences between the MOF and molecular crystal systems are relatively minor, the second-order SPC reveals clear distinctions. For the Cu(II) species, (1) exhibits markedly smaller coefficients, on the order of 10−9, compared to (3), which shows coefficients on the order of 10−5 for the gx and gy directions (Fig. 5). Similarly, for the vanadyl case (Fig. S27 and S28), the MOF (2) displays SPC coefficients on the order of 10−8, whereas (4) reaches values around 10−6 for gx and gy. Reflecting these differences in SPC magnitudes, the temperature-scaled SPC coefficients (Fig. S29–S32) also differs significantly between the systems, showing better agreement with experiments: (3) and (4) exhibit faster spin–lattice relaxation than (1) and (2). Fig. 4b shows the sum of the contributions to T1 from all vibrational modes scaled by the Bose–Einstein phonon occupation distribution. In this case, a better agreement was obtained with the experimental data that shows a less steep decrease in T1 for (1) compared to (3) than in the case of the first-order calculation. The improved agreement obtained with the second-order approach suggests that identifying the vibrational modes relevant to T1 at high temperatures – beyond those involved in the direct process – requires second-order spin–phonon coupling coefficients rather than only first-order terms.
![]() | ||
| Fig. 5 Second-order SPC coefficients for (a) Cu MOF (1) and (b) CuTPP (3). Vibrational modes were computed with VASP using the PBE-D3 functional. | ||
Although this second-order treatment improves the agreement with experiment, some discrepancies remain. These differences likely arise from methodological approximations, such as neglecting phonon dispersion over the entire Brillouin zone and omitting cross terms in the second-order coupling coefficients. While more rigorous treatments including these effects have been reported,11,73–75 they require computationally intensive methods that are beyond the scope of the present study. Nevertheless, the current approach captures the dominant experimental trends and provides valuable qualitative insight into the differences between the MOF and molecular crystal systems. Having established the overall trends in spin–phonon coupling and relaxation behaviour, we will now discuss the origin of the longer spin–lattice relaxation times observed in the MOF systems.
Animated visualization of the phonon modes in (1) shows that the lowest-frequency optical phonon mode at 18 cm−1 corresponds to a delocalized vibration involving collective motion of the entire MOF framework (Fig. 6). In this mode, the porphyrin units undergo rigid displacements while preserving their local ligand environments, resulting in only small spin–phonon coupling (Table S13; SPC coefficients: first-order ∼10−14, second-order ∼10−9). Although this phonon mode has a relatively small SPC coefficient (Fig. 5), its low energy leads to significant thermal population even at low temperatures, making it the dominant contributor to the spin–lattice relaxation of (1) via second-order processes (Fig. S29). However, its contribution to the T1 process remains much smaller than that of the most dominant phonon modes in (3) occurring at 208 cm−1 and 209 cm−1 due to the magnitude of their SPC couplings (on the order of 10−9 for (1) and 10−5 for (3) in gx and gy) (Table S15 and Fig. S29, S31). This difference in the size of spin–phonon coupling in (1) and (3) explains why the overall T1 process in (1) is slower than in (3), even though the lowest-frequency mode dominates in (1). This demonstrates that vibrational frequency shifts alone do not dictate relaxation behaviour, but the coupling strength of specific phonon modes is equally critical.
In the case of vanadyl species, the lowest experimental Raman bands in (2) and (4) are observed at similar frequencies (Fig. S33). For (2), the computed lowest-frequency optical phonon mode occurs at ∼20 cm−1, comparable to that of analogous (1) (Fig. S35). While the frequency of the lowest-optical phonon mode in (4) shows dependence on the DFT method, appearing at 19 cm−1 with PBE-D3/VTZP and 4 cm−1 with PBE-D3 (VASP), this lowest frequency mode consistently shows the largest contribution to the second-order process in (4), regardless of the computational method used. By contrast, although (2) also exhibits major contributions from very low-frequency modes in a similar range (Fig. S30; modes in the range between 16 cm−1 to 37 cm−1), the magnitude of the spin–phonon coupling remains smaller than (4), resulting in weaker spin–lattice relaxation contributions. Therefore, as in the Cu(II) case, the key factor distinguishing the MOFs and molecular crystal systems is not the phonon frequencies themselves, but rather the difference in the magnitude of the spin–phonon coupling.
By contrast, for the second-order spin–phonon coupling coefficients, the correlations between specific distortion types and coupling magnitudes is less clear than as the first-order SPC (Fig. 7 and Fig. S36–S38). Nevertheless, phonon modes that do not satisfy the symmetry requirements for first-order coupling (non-totally symmetric modes) tend to exhibit larger second-order SPC values. This trend is reasonable, as the first-order coupling requires the g-factor to vary linearly as a function of Q while the second-order coupling requires it to vary quadratically.69,83 While this may seem obvious, it highlights the need for caution for the common use of the first-order SPC coefficients to identify the vibrational modes that promote the spin–lattice relaxation based on the assumption that modes with large first-order SPC coefficients also have large second-order SPC coefficients. While there is a tendency for modes with small first-order SPC coefficients to exhibit larger second-order coefficients, this trend is less clear-cut than the prediction for first-order SPC from symmetry. For example, both A and B modes show similar sizes of second-order SPC in (1), as shown in Fig. 7a. As a result, predicting which vibrational modes play a significant role cannot be done as intuitively as for the first-order spin–phonon coupling, highlighting the importance of computational approaches in thoroughly understanding spin–lattice relaxation in addition to vibrational spectroscopy or group theory.
To examine this hypothesis, we quantified the in-plane and out-of-plane distortions in (1)–(4). For in-plane distortions, we evaluated the average M–N (M = Cu, V) bond-length change (defined as the average distance variation between the metal ion and the four coordinating nitrogen atoms) for each phonon mode. For out-of-plane distortions, we performed normal-coordinate structural decomposition (NSD) analysis of the porphyrin units. This analysis decomposes structural distortions into five fundamental modes: saddling (A2u), ruffling (B2u), doming (B1u), waving (Eg), and propelling (A1u) in an ideal D4h symmetric porphyrins (Fig. S39).50,51 The total out-of-plane distortion, Doop, can also be computed from this analysis. It should be noted that while these symmetry labels originate from the point group D4h, the actual phonon modes of the MOFs and molecular crystals are symmetry-lowered combinations of these distortions. Thus, the symmetry label in the NSD analysis do not directly correspond to the irreducible representations of the phonon modes. Each distorted structure at Q = 1 was analysed with respect to each phonon mode to identify correlations between specific distortion types, magnitude of the out-of-distortions and spin–phonon coupling strengths.
Fig. 8 shows the average distance between the Cu and the nitrogen and the absolute change in out-of-plane deformation Doop with respect to the equilibrium geometry (|ΔDoop|) obtained from the NSD analysis for (1) and (3). The results for (2) and (4) are shown in Fig. S40. When comparing the MOFs and molecular crystals, specifically (1) vs. (3) and (2) vs. (4), no drastic difference is observed in the distribution of M–N distances (Table S12). For Doop, (3) shows a smaller value than (1), while (4) shows a larger value than (2), showing no clear trend between the MOFs and molecular crystals (Table S12).
![]() | ||
| Fig. 8 The distribution of average Cu–N distance changes of phonon modes in (a) Cu MOF (1) (b) CuTPP (3) and out-of-plane distortion (|ΔDoop|) of phonon modes in (c) (1) (d) (3). | ||
Additionally, the correlation plots between the magnitudes of in-plane and out-of-plane distortions and their corresponding spin–phonon coupling coefficients (Fig. S41 and S42) are scattered, showing only a rather poor correlation for both the first-order and second-order SPC. This is in contrast with the generally accepted concept that, in order to achieve a longer T1, designing a rigid ligand that suppresses the ligand vibrations around the metal centre is the key. This absence of predictable patterns suggests that the spin–phonon coupling arises from more complex mechanisms that cannot be predicted by only the magnitude of distortions of the ligand around the metal centre. As noted in the previous section, symmetry is also likely to play a role in determining which mode can have large first-order and second-order SPC.
Consistent with this hypothesis, for the first-order spin–phonon coupling coefficients, a slightly better correlation was observed between specific distortion types and coupling strength (Fig. S43–S46). The distortion mode that shows a positive correlation is system-dependent. For (3), modes with larger saddling and ruffling distortions tend to show larger SPC coefficients (Fig. S45). By contrast, in (1), both ruffling and waving modes appear relevant (Fig. S43). In (4), phonon modes involving pronounced doming distortions tend to exhibit larger coupling coefficients (Fig. S46), while no clear correlation between specific distortion types and spin–phonon coupling was observed in (2) (Fig. S44). Although the symmetry of a given distortion cannot be directly assigned to a single irreducible representation of a phonon mode since phonon modes are mixtures of multiple distortion components, this system-dependent behaviour is in line with the idea that the symmetry imposes constraints on the spin–phonon coupling, as discussed in the previous section. For the second-order SPC (Fig. S47–S50), the correlation between the specific distortion mode and the size of the SPC coefficient is barely seen. This is similar to the symmetry analysis in the previous section, where the correlation between the symmetry and the size of SPC was not as obvious as the first-order SPC. This further suggests the more complicated nature of the second-order SPC and the difficulty in intuitively predicting the types of vibrational modes that have larger second-order SPC.
The distributions of the average distance change of the metal and the nitrogen (M–N; M = Cu or V) and ΔDoop as a function of phonon frequency are also shown in Fig. 9 for (1) and (3) and Fig. S51 for (2) and (4), and the overall patterns are quite similar across systems. In particular, the frequency-dependent profiles of M–N distance fluctuations are nearly identical between MOFs and molecular crystals. For instance, in the vanadyl systems (2) and (4), the phonon modes that perturb the M–N distance appear at nearly the same frequencies around 200, 280, 330, and 400 cm−1. In the Cu(II) systems, although (1) and (3) show slightly less similarity than the vanadyl analogues, the main Cu–N perturbing modes are still clustered around the same frequencies: approximately at 200, 280, 330, and 400 cm−1.
Despite the similarity of vibrational modes, the spin–phonon coupling (SPC) coefficients can differ significantly. For example, in compound (3), the dominant contributing modes are found at 208 and 209 cm−1 (PBE-D3, VASP; Fig. S31 and Table S15, B1, B2 mode respectively). A similar vibrational mode is present in (1) at a wavenumber of 197 cm−1 based on the NSD analysis. While this mode in (1) makes a significant contribution to the first-order SPC (Fig. S29, first-order SPC coefficients ∼10−6), its second-order contribution is much smaller (second-order SPC coefficients: ∼10−10) (Table S13). This difference can be partially attributed to the distinct symmetry or space group of the two systems. In the MOF (1), consistent with the group-theory based analysis, the 197 cm−1 mode is atotally symmetric A mode and induces a largely linear change in the g-factor as a function of Q, which effectively reduces the second-order coupling. Thus, the difference in spin–lattice relaxation behaviour between (1) and (3) can be partially explained by symmetry effects.
In the vanadyl systems, the second-order spin–phonon relaxation process in (4) is dominated by the lowest-frequency mode, located around 4 cm−1 (VASP, PBE-D3; Fig. S32). This mode exhibits largely phenyl rotations but is also accompanied by large out-of-plane distortions, as shown in Fig. S52. By contrast, while the lowest-frequency mode in (2) at 16 cm−1 also contributes significantly to the second-order process, primarily due to the temperature-dependent pre-factor, its actual SPC coefficient is much smaller (∼10−10) (Fig. S28 and Table S14). This difference is likely due to the significantly reduced ligand distortions around the vanadyl centre in (2), as illustrated in Fig. S51 and Table S14, S16. Therefore, in the case of vanadyl, the slower spin–lattice relaxation observed in the MOF (2) can be attributed to the suppression of low-frequency modes that strongly perturb the ligand environment. In (4), the molecular crystal supports a very low-frequency mode that is mainly characterized as phenyl rotation but also perturbs the ligand environment, enhancing relaxation. The MOF lattice modifies this mode and effectively suppresses the major relaxation pathway in (4). Although new low-frequency modes emerge in the MOF (2) due to the new framework structure, it perturbs the ligand environment much less, resulting in weaker spin–phonon coupling and longer T1.
Overall, while vibrational modes with large out-of-plane or in-plane distortions can exhibit strong first- or second-order spin–phonon coupling, as exemplified by the first optical phonon mode in (4), not all such modes behave this way. Instead, symmetry appears to play a critical role, and the magnitude of SPC coefficients is likely governed by a complex interplay of symmetry, distortion amplitude, and other factors. This indicates the difficulty in intuitively identifying the vibrational mode from commonly employed simple vibrational spectroscopy to suppress in order to achieve longer T1 and highlights the importance of computational approaches for accurately identifying the vibrational modes that strongly contribute to spin–lattice relaxation and for guiding the rational design of systems with longer relaxation times.
Our analysis revealed several important insights. First of all, first-order SPC calculations, which have been used to analyse the local mode impact on T1,13,16,20,39,41,69–71 did not reproduce the experimentally observed differences in T1, emphasizing the need to consider second-order SPC coefficients. In particular, modes that are symmetry-forbidden for first-order coupling often exhibit significant second-order contributions. Secondly, our results show that the frequency alone does not explain T1 differences. While symmetry seems to play a somewhat important role in determining the size of SPC both for the first-order and second-order, the correlation with the magnitude of the SPC coefficients and symmetry is not straightforward, especially for the second-order SPC. Predicting which vibrational modes dominate spin–lattice relaxation based solely on symmetry remains challenging. We also analysed the correlation between porphyrin ligand distortions and SPC coefficients. While modes involving large out-of-plane or in-plane distortions can yield strong SPC, not all do. The presence of such distortions alone does not guarantee strong coupling, and the contribution of each mode appears to depend on a complex interplay between symmetry, distortion amplitude, and additional factors that were not included in the present analysis.
Overall, our results demonstrate that no single descriptor, neither frequency, symmetry, nor distortion magnitude, can reliably predict the spin–phonon coupling strength. Particularly for second-order processes, these contributions seem to arise from complex mechanisms that cannot be captured through intuitive analysis alone. Therefore, while experimental characterization such as low-frequency vibrational spectroscopy is useful, computational approaches are also essential for identifying key vibrational modes responsible for spin–lattice relaxation and for guiding the rational design of ligands and frameworks that support longer T1 times in molecular spin qubits.
Supplementary information (SI) containing additional experimental details, figures, and analysis is available. See DOI: https://doi.org/10.1039/d5cp04453g.
| This journal is © the Owner Societies 2026 |