Cristian
Villa-Pérez
*a,
Andoni
Zabala-Lekuona
*b,
Iñigo J.
Vitorica-Yrezabal
c,
José Manuel
Seco
b,
Javier
Cepeda
b,
Gustavo Alberto
Echeverría
d and
Delia Beatriz
Soria
*a
aCEQUINOR (CONICET, CCT – La Plata), Departamento de Química, Facultad de Ciencias Exactas, Universidad Nacional de la Plata, Bv. 120 no. 1465, 1900, La Plata, Argentina. E-mail: cristianvilla@quimica.unlp.edu.ar; soria@quimica.unlp.edu.ar
bDepartamento de Química Aplicada, Facultad de Química, Universidad del País Vasco/Euskal Herriko Unibertsitatea (UPV/EHU), Paseo Manuel Lardizabal no. 3, 20018, Donostia, Spain. E-mail: andoni.zabala@ehu.eus
cDepartamento de Química Inorgánica, Facultad de Ciencias, Universidad de Granada, 18071, Granada, Spain
dIFLP (CONICET, CCT – La Plata), Departamento de Física, Facultad de Ciencias Exactas, Universidad Nacional de la Plata, 47 y 115, 1900, La Plata, Argentina
First published on 9th January 2024
Monomeric [Co(SDZ)2phen] (1) and [Co(SDZ)(bq)Cl] (2) complexes (SDZ = sulfadiazine, phen = 1,10-phenanthroline, and bq = 2,2′-biquinoline) have been synthesized and characterized. X-ray diffraction studies indicate that SDZ acts as a bidentate ligand coordinating through the sulfonamide and the pyrimidine N atoms in both compounds. In complex 1, the coordination sphere consists of two SDZ ligands and a bis-chelating phen ligand, giving rise to a CoN6 coordination sphere. On the other hand, 2 has a CoN4Cl core, with two N-atoms from SDZ and two from the bq ligand. Both compounds have been studied by dc and ac magnetometry and shown to display slow magnetic relaxation under an optimum external dc field (1 kOe) at low temperatures. Moreover, compound 2 displays long range magnetic ordering provided by spin-canted antiferromagnetism, which has been characterized by further field-dependent magnetic susceptibility measurements, FC/ZFC curves, hysteresis loops and frequency-independent ac curves. The signs of the calculated D parameters, positive in 1 and negative in 2, have been rationalized according to the two lowest-lying transitions in the orbital energy diagrams derived from ab initio ligand field theory (AILFT). In a subsequent attempt to reveal the possible hidden zero-field SMM behaviour, Ni(II)-based 3 and Co(II)-doped Ni(II)-based (with a Ni:
Co ratio of 0.9
:
0.1) heterometallic compound 2Ni were synthesized.
Co(II)-based complexes are good candidates for the development of SMMs due to their large magnetic anisotropy, which originates from a strong contribution of first-order spin–orbit coupling to the total magnetic moment. The magnetic properties of these complexes are strongly influenced by the coordination environment provided by the ligands, and there is a great interest in developing complexes with coordination numbers as low as possible, due to their expected large magnetic anisotropy.17 High coordination numbers are known to suppress the orbital contribution (L, orbital angular momentum) and subsequent magnetic anisotropy. Indeed, the ground-state could be appropriately described by the spin (S) term. When low coordination numbers are obtained, the d orbitals fall within narrower energy ranges, simulating 4f orbitals of lanthanide compounds. For instance, the current and previous records of the largest barriers to the reversal of magnetization in transition metal complexes are two linear Co(II) complexes.17,18 Nevertheless, it has also been proved that compounds with higher coordination numbers could have significant magnetic anisotropy. Several Co(II) single-ion magnets with coordination numbers ranging from 2 to 8 and with diverse geometries have been reported so far.17–20,22–27 Hence, selecting appropriate ligands for the development of suitable ligand fields is fundamental, as they will modulate the magnetic anisotropy of the metal ion and, therefore, the potential SMM/SIM behaviour.
In this work, we report novel complexes that are derivatives of the sulfadiazine ligand, which is a widely used antibiotic in both human and veterinary medicine. In addition, it has several coordination modes that could lead to a wide variety of materials.28 Note that we planned this research project as a continuation of a previously published work reported by some of us, in which two sulfadiazine six-coordinated Co(II) complexes with the ancillary ligands 2,2′-bipyridine and 6-methoxyquinoline were described.22 Both compounds had been shown to be field-induced SMMs with Ueff values of 50.6 K (2,2′-bipyridine derivative) and 13.7 K (6-methoxyquinoline derivative). In the present case, the heterocyclic compounds 1,10′-phenanthroline and 2,2′-biquinoline have been used as ancillary ligands. The former one could provide a similar structure to the one studied with 2,2′-bipyridine, while the latter one was selected as a bulkier ligand with the aim of sterically hindering some coordinating positions and, thus, for obtaining lower coordination numbers.
The present work covers the synthesis and characterization of three novel compounds, followed by experimental and theoretical magnetic studies.
The crystal data and refinement results are summarized in Table S1.† CCDC 2040162, 2040161 and 2309444 for 1–3,† respectively, contain the supplementary crystallographic data for this paper.
The X-ray powder diffraction (XRPD) patterns were recorded over powdered samples (Fig. S1 and S2†). For data acquisition, a Philips X'PERT powder diffractometer was used with Cu-Kα radiation (λ = 1.5418 Å) covering the range of 5 < 2θ < 50° with a step size of 0.026° and an acquisition time of 2.5 s per step at 25 °C.
The coordination geometry distortion can be evidenced by the bond distances and angles in the coordination spheres (see Table S4†). For 1, the coordination sphere around the Co(II) ion consists of the phen nitrogen atoms at 2.102(7) and 2.113(5) Å and four N-atoms from two SDZ molecules at 2.106(4), 2.193(5), 2.276(5) and 2.094(5) Å, completing a distorted octahedral geometry. All bonding angles deviate from the ideal values. For instance, both SDZ ligands bond to the metal ion through the sulfonamido and one pyrimidine nitrogen atoms, establishing four-membered CoNCN′ rings with N–Co–N′ chelating angles of 60.88° and 62.48°. On the other hand, the phen ligand chelates the metal ion forming an angle of 78.41°.
In 2, the cation is surrounded by a distorted square pyramidal environment. The equatorial positions are occupied by a chlorine atom at 2.2746(2) Å, two nitrogen atoms from an SDZ ligand at 2.1166(2) and 2.1735(2) Å, and another one from bq at 2.061(4) Å. The apical position of the pyramid is occupied by the second N-atom of the bq at 2.039(5) Å. Co(II) is located 0.405 Å above the base of the distorted square pyramid. The angles in the pyramidal base are deviated from the ideal 90°, with values ranging from 61.87° (N16–Co–N17) to 99.43° (Cl1–Co–N16). Similarly, the angles between the base atoms and the apical N220 deviate from the ideal value, with values as low as 79.67° (N21–Co–N220) and as high as 119.01° (N16–Co–N220).
Furthermore, the crystal lattices are stabilized because of the presence of several intermolecular hydrogen bonds (Table S5†). In both complexes, the main intermolecular interactions, which stabilize the crystal structure, are the N–H⋯O hydrogen bonds connecting the anilinic SDZ nitrogen atom to the sulfonamide oxygen atoms, generating an extended structural pattern. In addition, π⋯π intermolecular interactions also play an important role in the molecular packing of compound 2 as detailed below (Fig. S3†).
In 1, the amino nitrogen and oxygen atoms of neighboring SDZ ligands, symmetry-related by the c and a-glide planes, are involved in four N–H⋯O hydrogen bonds, forming a 3D N–H⋯O hydrogen bonding network that is spread along the three crystallographic directions (Fig. 2). In addition, the supramolecular crystal building is further stabilized by weaker C–H⋯N and C–H⋯O hydrogen bonds involving carbon atoms of the phenylamine group and pyrimidine nitrogen or sulfonamide oxygen atoms. Among all intermolecular pathways, the shortest Co⋯Co distances are of 8.773(1) Å in 1.
In 2, the crystal building involves various types of interactions of variable nature and strength that set the neighbouring complexes at different distances between Co(II) centres. The main interaction between the complexes is due to the formation of dimeric units through relatively strong π⋯π interactions (given the large overlap between the central aromatic rings of bq ligands with C⋯C distances in the 3.5–3.7 Å range; see Table S6† for further details) and weaker C211–H211⋯Cl1 hydrogen bonds. This can be denoted as a centrosymmetric interaction pathway due to the presence of an inversion center that places the stacked bq ligands at ca. 3.57 Å between the centroids of interacting aromatic rings (see Fig. S3†). Interestingly, within these layers, bq ligands are packed in a fashion that resembles the herringbone packing of polycyclic aromatic hydrocarbons.51,52 As a consequence, the Co(II) ions of neighbouring complexes are situated at a distance of 7.2 Å, which is the shortest intermolecular pathway connecting the metal atoms. Moreover, each of the complexes of the centrosymmetric dimers establishes C–H⋯π interactions among SDZ ligands in addition to C221–H221⋯Cl1 (involving the aromatic bq carbon atom) weaker bonds to generate infinite arrays of complexes along the crystallographic a axis, resulting in complexes being separated by a Co⋯Co distance of 9.3 Å. As a result of both interaction pathways, 2D layers arranged along the (101) plane are formed (Fig. 3). Additionally, note that arrays of π⋯π stacked dimers are further linked with each other through additional π⋯π interactions between the peripheral aromatic rings of bq ligands (Fig. S3†). Finally, the 2D layers are further piled up along the crystallographic b axis in such a way that dimeric entities are displayed alternately with two alternative orientations with respect to the piling direction, given that neighbouring layers are related by the glide n plane. As a result, two inequivalent additional intermolecular pathways are observed between complexes pertaining to alternate layers: (i) a non-symmetric pathway involving the N117–H117A⋯O110 hydrogen bond established between the amino nitrogen and sulfonamido oxygen atoms of two neighboring SDZ (imposing a Co⋯Co distance of 11.2 Å) and (ii) a non-symmetric pathway along the N117–H117B⋯Cl1 hydrogen bond established by the amino nitrogen atom of the SDZ ligand (imposing a Co⋯Co distance of 11.9 Å). Most importantly, these two superexchange pathways share the absence of a symmetry element between the complexes, and their relative orientation obeys to restrictions of the overall packing, in such a way that pyramidal environments are relatively twisted showing non-fully parallel nor anti-parallel orientations (to give an orientative measure, angles between Cu–Napical vectors are of 59.3° and 82.3° for the non-symmetric pathways 1 and 2, respectively).
For 1, the room temperature χMT product of 2.84 cm3 mol−1 K is significantly higher than the expected spin-only value for an octahedral Co(II) ion (1.87 cm3 mol−1 K with g = 2.01), which suggests the presence of certain spin–orbit coupling (Fig. 4). On cooling down, the χMT value remains almost constant before a final and more abrupt drop below 75 K, reaching a minimum value of 1.68 cm3 mol−1 K at 2 K. This progressive decrease may be attributed to the first-order SOC effect as sizeable antiferromagnetic interactions have been ruled out in view of the long distances between spin carriers in the structure (the shortest interactions impose Co⋯Co distances of ca. 8.7 Å) and, most importantly, the absence of remarkable π–π or hydrogen bonding interactions to mediate magnetic exchange. On account of the CoN6 distorted octahedron present in compound 1, the potential magnetic anisotropy usually observed in these metal centres53 was further corroborated by isothermal magnetization curves collected in the 2–5 K range (Fig. 4, inset), as they do not reach the theoretical saturation for an S = 3/2 system (Msat = 3.3μB, with g = 2.2).
![]() | ||
Fig. 4 Variable-temperature dc magnetic susceptibility data for 1 collected under a 1 kOe applied dc field. Inset: The variable-field magnetization curves recorded in the 2–5 K temperature range. The continuous lines in both plots represent fits to eqn (1) using the PHI program. |
In order to evaluate the sign and magnitude of the zfs parameter, we simultaneously fitted both the susceptibility and magnetization data with the spin Hamiltonian shown in eqn (1) using the PHI program:54
![]() | (1) |
The temperature dependence of the χMT product for compound 2 notably differs from 1, especially at the lowest temperatures (Fig. 5). At room temperature, the χMT value of 3.41 cm3 mol−1 K is much higher than the expected spin-only value for a Co(II) ion. On cooling down, this value smoothly decreases, reaching a minimum value of 3.17 cm3 mol−1 K at 46 K, and then abruptly increases up to 6.34 cm3 mol−1 K, describing a maximum at 11 K, eventually dropping to 2.26 cm3 mol−1 K at 2 K. This behaviour is indicative of a weak net ferromagnetic ordering in the compound, which is quite surprising in view of the molecular nature of the compound composed of isolated complexes. Nonetheless, the relatively short intermolecular Co⋯Co distances imposed by π⋯ π interactions between 2,2′-biquinoline ligands combined with other weak couplings along non-symmetric pathways in the structure seem to be responsible for the observed long-range ferromagnetic coupling (see Fig. S5†).
![]() | ||
Fig. 5 Variable-temperature dc magnetic susceptibility data for 2 collected under a 1 kOe applied dc field. Inset: χMT plots recorded under different external magnetic fields. |
The occurrence of canted antiferromagnetism was further confirmed by several additional measurements. On the one hand, the susceptibility curves measured at variable fields reveal a strong field-dependent response (Fig. 5, inset), which agrees with the usual behaviour observed for ferromagnetic compounds. Moreover, field-cooled (FC) and zero-field-cooled (ZFC) χM curves show a clear bifurcation at 11 K (Fig. 6, top), which fits with the temperature at which frequency-independent maxima are observed in the ac dynamic magnetic measurements (vide infra). Lastly, isothermal magnetic hysteresis loops were recorded in the 2–5 K temperature range, displaying a noticeable opening at all studied temperatures (Fig. 6, bottom). The S-shaped curves observed at the highest measured temperatures (4–5 K) are specially worth mentioning, where a large positive slope at almost zero applied fields (Hdc < 0.1 T) is followed by a lower slope at higher fields. This shape, quite smoothed in the present case due to the weak magnetic interactions (vide infra), is indicative of a change in the regime of the magnetic ordering, in line with canted ferromagnetic behaviour. Furthermore, at 2 K remnant magnetization and coercive field values of 0.06μB and 315 Oe were measured, respectively. According to eqn (2), a small canting angle of 1.44° was estimated (MR and MS stand for the remnant and saturation magnetization values, respectively):
![]() | (2) |
![]() | ||
Fig. 6 For 2, temperature-dependent ZFC and FC molar susceptibility curves in the low temperature region (top) and magnetic hysteresis loops recorded in the 2–5 K temperature range (bottom). |
Considering that complex 2 does not crystallize in a non-centrosymmetric space group, we assume that the spin-canting behaviour arises from the single-ion magnetic anisotropy of Co(II) ions and the relative dispositions between adjacent complexes in the structure, forced by intermolecular Co⋯Co interactions, although there could be some superexchange pathway allowing for substantial antisymmetric exchange. In particular, the present structure is characterized by multiple intermolecular pathways along the spin carriers (take into account that intermolecular pathways in Fig. 3 represent all potential magnetic couplings), all of which are expected to result in very weak exchange interactions. Note, in this sense, that no reliable calculations could be performed to provide the magnitude of the exchange coupling through the exchange pathways owing to the limitations of DFT to estimate the value of J constant in these kinds of superexchange pathways.60 In principle, the strongest magnetic interaction should come from the so-called intralayer centrosymmetric pathway occurring through the π⋯π stacking between bq ligands because (i) it brings the shortest Co⋯Co intermolecular distance (7.2 Å) in the structure and (ii) the spin density is relatively large over the bq ligand (Fig. S5†). As reported in previous works, π–π interactions are known to yield weak ferromagnetic couplings,61,62 which seems to be the case of the present compound in view of the structural characteristics of the interacting bq ligands (see Table S6†). Assuming that this interaction is the strongest one along the 2D planes (drawn in red and green in Fig. 3), there would be a net ferromagnetic ordering in the layers given the weaker superexchange couplings occurring through the remaining bridges within the layers. Hereafter, assuming that weak antiferromagnetic interactions may occur along the interlayer non-symmetric pathways (involving Co⋯Co of ca. 11 Å, see bottom of Fig. 3), the magnetic moment of the 2D layers should not be fully cancelled with each other when packed along the crystallographic b axis. In fact, a representation of the D-tensor frame in two Co centres interacting through this large pathway shows how magnetic axes are not fully parallel, but canted to each other, and thus, given the absence of symmetry elements different from n glide planes, uncompensated magnetic moments along the packing direction (b axis) lead to the observed spin-canting behavior (Fig. S9†). Taking into account that this explanation cannot be fully supported by calculations (in view of the impossibility to reliably calculate J constants), another possible explanation would suppose that interactions along the interlayer non-symmetric pathways are the main superexchange bridges governing the magnetic properties (canting the spins between individual complexes at low temperature), which is a priori more difficult to understand in view of the long Co⋯Co distances and low spin density present over the interacting atoms. In any case, the presence of the observed canting (involving S-shaped hysteresis curves) is well explained according to these two exchange pathways containing Co ions related by n glide planes, in which antisymmetric exchange could be non-zero. As an illustrative example, we would like to cite the research work reported by Zhang et al., in which they compared several one-dimensional Co(II) based compounds.63 As they state, the coordination spheres in all compounds are comparable, as well as the intrachain magnetic interactions. However, different hydrogen-bond mediated Co⋯Co distances are responsible for activating/deactivating magnetic ordering in addition to the occurrence of SCM (single-chain magnet) behaviour.
Taking into account that the long-range magnetic ordering prevents us from fitting the dc magnetic data of complex 2 and estimating the zfs parameters, theoretical calculations are essential in their evaluation. In this case, the calculation gives a large negative D = −59.4 cm−1 along with a large rhombicity parameter (E/D = 0.21), which makes the sign of the main D parameter meaningless (see Table S7†), as discussed before for compound 1.55–57 In order to understand the origin of the zfs parameters in this compound, we examined the electronic configurations for the active space of the ground and lowest-lying states (see Table S10†). As expected for a severely distorted square pyramid, all d-orbitals are separated in energy and the (dyz)2(dxz)2(dxy)1(dz2)1(dx2−y2)1 configuration better represents the ground state according to the ab initio ligand field theory (AILFT) method (Fig. 7). It is worth noting that dxy and dxz orbitals cannot be distinguished and appear to be admixed between the HOMO and LUMO orbitals as a consequence of the large distortion present (Table S9†), which could explain the sign of the lowest-lying transitions involved in the zfs parameters of the compound. As observed in Table S7,† computational calculations predict negative and positive signs for the first and second excitations, which mainly represent the sign and magnitude of the D parameter. Therefore, according to the different |ml| values of the dxy (±2) and dxz/dyz (±1) orbitals and the transition energies (Table S10†), the main dxz → dxy transition should govern the first transition, whereas the second one should mainly consist of the dyz → dxz (taking into account the minor, but still significant, contribution of dxz to the third orbital) transition, which can only be explained according to the previous admixture. Moreover, such an admixture involving the dxy and dxz orbitals, with a predominant equatorial and axial component, respectively, could also be the origin of the large rhombicity present in the compound and the ambiguous overall sign of the zfs parameters.
![]() | ||
Fig. 7 NEVPT2-AILFT computed d-orbital splitting for compound 2. The dashed lines represent the first and second excitations, which contribute to the zfs parameters of the compound. |
In spite of the significant magnetic anisotropy present, the occurrence of zero-field SMM behaviour is not expected in the present case in view of the large value for the matrix element (0.5μB) connecting both ground Kramers doublets (Fig. S11†), which suggests that a considerable tunnelling effect could completely quench the slow magnetic relaxation. In addition, the possible SIM behaviour in this compound would compete with the long range magnetic ordering provided by spin-canting, which is the reason for subsequent dilution attempts.
As described in the upcoming sections, the Ni(II) based counterpart 3 was synthesized in order to simulate a diamagnetic dilution of 2 because it was not possible to synthesize the isostructural Zn(II) counterpart. This unconventional strategy is based on a recently published work reported by Zadrozny and coworkers,64 where they took advantage of the positive D value of the Ni(II) counterpart in (Ph4P)2[M(SPh)4] (where M is Mn(II), Fe(II), Co(II) or Ni(II) in their study) that simulates a diamagnetic ground state (MS = 0). Due to the lower expected magnetic anisotropy, the χMT curve of 3 does not show any canted antiferromagnetism (Fig. S12†). In fact, this curve can be properly fitted with the PHI software by using the following Hamiltonian:
![]() | (3) |
As observed in Fig. S12,† the susceptibility curve fits well for both positive and negative D values. In the first attempt, the curve was properly fitted, affording D = +12.5(2) cm−1, g = 2.2(0) and R = 2.2 × 10−2, but without including the intermolecular −zJ interactions. In the second attempt, we obtained a similar result with the following set of parameters: D = −12.4(9) cm−1, g = 2.3(0) and R = 7.1 × 10−2, including zJ = −0.5 cm−1. The origin of such ambiguity is also explained by the large rhombicity present in the compound, as suggested by CAS-SCF/NEVPT2 calculations (vide infra, see Table S8†). Considering the short M⋯M distances found in the crystal structure and that intermolecular interactions govern the static and dynamic (vide infra) properties of the Co(II) based counterpart, both fitting procedures seem reasonable and, therefore, the ground state of 3 could be either MS = ±1 or MS = 0. Accordingly, the Ni/Co heterometallic mixture was studied to explore the possible improvement of SMM behaviour by means of the dilution of Co in a Ni based matrix.
At a zero applied dc field, no signal was observed in the χ′′M(T) plot for 1, which may be due to the existence of a fast QTM process hiding the desired SMM behaviour (Fig. S13†). In view of that, the field-dependent ac response was studied at 2 K. As observed in Fig. S14 and S15†, the SMM behaviour arises even at the lowest applied dc field of 250 Oe. Notably, the slowest relaxation time was found at 1 kOe, becoming faster at 2.5 kOe due to the enhancement of a field-induced direct process. Thus, temperature- and frequency-dependent measurements were carried out with an external field of 1 kOe (Fig. 8). Under these conditions, QTM is at least partially quenched and 1 displays temperature- and frequency-dependent maxima below 4 K. Both χ′′M(χ′M) or Cole–Cole plots and χ′′M(ν) curves were fitted to the generalized Debye model within in the 2–4 K temperature range (Fig. S17 and S18†). As expected for an octahedral Co(II) complex, the relaxation of magnetization is best described by a Raman process (Fig. 8, top, inset) instead of an Orbach mechanism involving excited states. In fact, a linear fit of the Arrhenius plot involves Ueff = 9.2(2) K (6.4 cm−1), a value that is much lower than the calculated 2D or the theoretically calculated energy gap between the ground and excited Kramers doublets by means of SINGLE_ANISO (244 K, 169.5 cm−1; see Fig. S10†). In fact, a single Raman mechanism well describes the temperature dependence of the relaxation times, in agreement with the low α values calculated from the Cole–Cole plots (Fig. S17†). Thus, the temperature-dependence of the relaxation times was fitted to the following eqn (4):
τ−1 = BTn | (4) |
The best fit provided B and n values of 2055(67) s−1 K−n and 2.36(2), respectively. Additionally, the lack of zero-field SMM behaviour is also well explained by the matrix element within the ground state, which predicts a large tunnelling phenomenon with a value of 1.7μB (see Fig. S10†).
In good agreement with the long-range weak ferromagnetic ordering present in 2, the data collected at zero-field showed temperature- and frequency-independent maxima in both of the χ′M(T) and χ′′M(T) signals (Fig. S19† and Fig. 8, middle). The fact of having these maxima at 11 K is in good agreement with the bifurcation temperature found in the FC/ZFC measurement. In order to confirm that the maxima do not correspond to the SMM behaviour, the Mydosh parameter,65ϕ, was calculated using the formula ϕ = ((ΔTP/TP)/Δlogf), giving a value of ϕ = 0.04, which is consistent with a glassy state probably derived from the opposed weak ferro/antiferromagnetic interactions and not SMM behaviour.
Without discarding the possibility of a hidden SMM behaviour, field-dependent measurements were performed at 2.8 K for 2. As depicted in Fig. S20 and S21,† an optimum field of 1 kOe was determined for complex 2. Under these optimal conditions, frequency-dependent maxima were measured below 6 K within χ′′M(T) plots in agreement with a field-induced SMM behaviour. Cole–Cole and χ′′M(ν) plots were fitted within the 3.0–5.8 K temperature range and, once again, relaxation times were fitted to a Raman mechanism. The fitting to eqn (4) provided B and n values of 0.58(2) s−1 K−n and 6.58(2), respectively. However, taking into account that the α values extracted from the Cole–Cole plots are slightly larger than those calculated for 1 (Fig. S23†) and considering the calculated D value, we also considered the simultaneous occurrence of Orbach and Raman mechanisms using eqn (5):
τ−1 = τ0−1![]() | (5) |
Unfortunately, reasonable fitting parameters were not obtained. Before totally discarding the Orbach mechanism, we tried other attempts involving both QTM and direct processes along the Orbach relaxation pathway, but all our attempts were unsuccessful.
In view of the non-negligible Co⋯Co interactions appearing in 2, we made several attempts in order to obtain the isostructural and diamagnetic Zn(II) counterpart to study the relaxation behaviour of a single [Co(SDZ)(bq)Cl] molecule in a diluted Zn-based matrix. However, no attempts involving ZnCl2 afforded the desired compound. As an alternative route, we synthesized the Ni(II) counterpart 3. Very recently, Zadrozny and coauthors have reported the relaxation dynamics of the (Ph4P)2[Co(SPh)4] zero-field SMM diluted in different paramagnetic matrices. As they show, when the SMM is diluted in a Ni(II) matrix with D > 0, the MS = 0 sublevel is the only populated one at low temperatures with no spin angular momentum, which simulates, somehow, a dilution in a diamagnetic matrix alternative to the use of Zn(II). In this case, the dilution favours slower relaxation times due to an effective quenching of QTM.
In our work, we have tried to mimic their strategy for our compound 2. Hence, compound 2Ni was successfully synthesized by using a 1:
10 Co
:
Ni ratio, showing purity and homogeneity by PXRD analysis (Fig. S2†). Once again, dynamic ac magnetic measurements were initially performed under a zero applied dc field. Considering that 3 does not show any measurable spin-canted effect in the χMT curve, we did not expect it within 2Ni. However, as clearly observed in Fig. S25 and S26,† this diluted analogue displays frequency-independent maxima at 6.5 K in agreement with the long-range magnetic ordering shown by the cobalt analogue. In view of this, and with the aim of comparing the relaxation times measured in the same experimental conditions, ac magnetic properties of 2Ni were studied under an external magnetic dc field of 1 kOe. The χ′′M(T) plots reveal field-induced SMM behaviour with maxima below 5.2 K, a slightly lower temperature than that for 2 (Fig. S28†). Relaxation times and α values were obtained by fitting the Cole–Cole and χ′′M(ν) plots in the 2.0–5.2 K temperature range. As observed in Fig. 9, relaxation times for 2Ni display a larger curvature than that expected for a single Raman mechanism (dashed blue line). This is supported by the slightly larger α values obtained for this compound, which suggest the occurrence of an additional magnetic relaxation pathway. Thus, we considered the simultaneous presence of a Raman and a direct process by the following equation:
τ−1 = BTn + AdirectT | (6) |
![]() | ||
Fig. 9 Comparison of the Arrhenius plots for the relaxation times obtained for compounds 2 (red) and 2Ni (blue) measured under the same experimental conditions. |
The fit afforded the following set of parameters: B = 27(6) s−1 K−n; n = 4.4(1) and Adirect = 1881(120) s−1 K−1. Note that a similar fit could be obtained by replacing the direct mechanism with QTM (Fig. S31†). In any case, the most important conclusion is that relaxation times are faster for 2Ni than for 2 in the whole temperature regime.
It is clear that this system is not appropriate to perform the mentioned approach. Indeed, in view of the evolution of the relaxation times, we assume that the Ni(II) based counterpart must have a non-desired negative D value with MS = ±1 as the ground state. Ab initio calculations based on an optimized nickel-based model give D = −21.5 cm−1 although the negative sign is, again, not meaningful in view of the large rhombicity (E/D = 0.22, see Table S8†). Consequently, these results could explain why the relaxation times are not slowed down in this case because, although those few [Co(SDZ)(bq)Cl] complexes were aimed to be surrounded by either completely diamagnetic (with Zn(II)) or somewhat diamagnetic ions at low temperatures (Ni(II) with D > 0), they possess neighbouring paramagnetic molecules. In this sense, it is worth mentioning that the system that we have studied in this work and (Ph4P)2[Co(SPh)4] display notable differences. On the one hand, in the previously reported case, the paramagnetic centres within the crystal structure are much farther from each other and no sizeable interaction is observed. In our case, instead, the short Co⋯Co distances appeared to be fundamental when explaining the magnetic properties. On the other hand, (Ph4P)2[Co(SPh)4] behaves as a zero-field SMM, while in our case an external magnetic field needs to be applied to observe slow relaxation of magnetization.
Footnote |
† Electronic supplementary information (ESI) available. CCDC 2040162 (1), 2040161 (2) and 2309444 (3). For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d3dt02359a |
This journal is © The Royal Society of Chemistry 2024 |