Xiu-Fang Ma‡
a,
Xin-Lan Hou‡b,
Ye-Hui Qina,
Qian Tenga,
Song-Song Bao
a,
Yu-Xi Tian
*b and
Li-Min Zheng
*a
aState Key Laboratory of Coordination Chemistry, School of Chemistry and Chemical Engineering, Collaborative Innovation Center of Advanced Microstructures, Nanjing University, Nanjing 210023, China. E-mail: lmzheng@nju.edu.cn
bState Key Laboratory of Analytical Chemistry for Life Science, Key Laboratory of Mesoscopic Chemistry of MOE, School of Chemistry and Chemical Engineering, Nanjing University, Nanjing 210023, China. E-mail: tyx@nju.edu.cn
First published on 22nd April 2025
Stimuli-responsive lanthanide-based single-molecule magnets (Ln-SMMs) are attractive for their potential in information storage, sensors and molecular devices. However, challenges remain in synergistically tuning the magnetic and optical properties of Ln-SMMs at the same temperature. Herein, we report a mononuclear dysprosium complex containing anthracene units, namely [Dy(SeCN)2(NO3)(depma)2(4-hpy)2] (1DySeCN) (depma = 9-diethylphosphonomethylanthracene, 4-hpy = 4-hydroxypyridine). It shows a thermally induced phase transition attributed to an order–disorder transition of the axial 4-hpy ligand, and a thermochromism due to a slight slipping of the anthracene pairs. Compound 1DySeCN can undergo [4 + 4] photocycloaddition reaction in a single-crystal-to-single-crystal (SC–SC) transformation manner at room temperature to form the 1D coordination polymer [Dy(SeCN)2(NO3)(depma2)(4-hpy)2]n (2DySeCN), accompanied by a luminescence switch from yellow to blue-white. Meanwhile, significant changes in SMM properties are observed with the reduction of energy barrier from 334 K to 144 K and the narrowing of the butterfly-like hysteresis loop. We further investigated the effect of temperature on the photodimerization reaction of anthracene in 1DySeCN, and found that the compound can still undergo efficient photodimerization at temperatures as low as 200 K. At this temperature, the magnetic susceptibility (χMT) value of the dilute compound 1DySeCN@Y also changed significantly before and after light irradiation. This study provides the first example of lanthanide–anthracene compounds synergistically modulating the magnetic and luminescent properties at the same temperature.
To address these questions, we herein report a new Dy–anthracene complex, [Dy(SeCN)2(NO3)(depma)2(4-hpy)2] (1DySeCN, 4-hpy = 4-hydroxypyridine, depma = 9-diethylphosphonomethylanthracene, Scheme 1), and its photophysical, photochemical and magnetic properties at and below room temperature. 1DySeCN was designed and synthesized because of its structural similarity to the previously reported compound [Dy(SCN)2(NO3)(depma)2(4-hpy)2] (DySCN, Scheme 1)31 except for the replacement of SCN− with SeCN−. We envisioned that the incorporation of SeCN− would lead to a weakening of the Dy–N bond at the equatorial position, which in turn would enhance the magnetic anisotropy. Indeed, 1DySeCN shows SMM behaviour at zero dc field with an effective energy barrier (Ueff) of 334 K, which is higher than that of DySCN (Ueff = 277 K). The dilute sample 1DySeCN@Y displays an open butterfly-shaped magnetic hysteresis below the blocking temperature (TB = 4.2 K) with a coercivity (Hc) of 1200 Oe at 2.0 K. To the best of our knowledge, the Ueff, TB and Hc values are the highest among the known photoresponsive luminescent Ln-SMMs. In addition, 1DySeCN undergoes photocycloaddition reaction in a single-crystal-to-single-crystal (SC–SC) manner at room temperature to form a 1D chain compound [Dy(SeCN)2(NO3)(depma2)(4-hpy)2]n (2DySeCN, depma2 is the photo-dimerized depma), accompanied by remarkable changes in photoluminescent (PL) and magnetic properties. Interestingly, the photocycloaddition reaction of 1DySeCN can occur at temperatures down to 140 K. The in situ PL measurements on a single crystal of 1DySeCN revealed a significant photo-induced change in luminescence when the temperature was as low as 200 K. Impressively, the dilute sample 1DySeCN@Y showed photo-triggered luminescence and magnetic changes at 200 K. This work opens up the possibility of synergistically modulating the magnetic and luminescent properties of Ln-SMMs by photocycloaddition at the same temperature.
Complex | 1DySeCN-293 | 1DySeCN-193 | 1DySeCN-130 | 2DySeCN-293 | 2DySeCN-193 |
---|---|---|---|---|---|
Temperature | 293 K | 193 K | 130 K | 293 K | 193 K |
Crystal system | Monoclinic | Triclinic | Triclinic | Monoclinic | Monoclinic |
Space group | P21/m | P![]() |
P![]() |
P21/m | P21/m |
a (Å) | 9.7539(12) | 9.6729(5) | 9.7055(12) | 9.585(9) | 9.5847(12) |
b (Å) | 25.787(3) | 11.1667(6) | 11.1138(13) | 25.97(2) | 25.809(3) |
c (Å) | 11.2182(14) | 25.6156(16) | 25.416(3) | 11.340(12) | 11.2751(13) |
α (0) | 90 | 90.315(2) | 90.469(3) | 90 | 90 |
β (0) | 111.587(4) | 92.411(2) | 92.518(4) | 112.620(19) | 112.946(3) |
γ (0) | 90 | 112.001(2) | 112.134(3) | 90 | 90 |
V (Å3) | 2623.8(5) | 2562.4(3) | 2536.1(5) | 2606(4) | 2568.4(6) |
Dy–O (axial)/Å | 2.208/2.231 | 2.228/2.240 | 2.227/2.243 | 2.250/2.222 | 2.219/2.225 |
Dy–O (equatorial)/Å | 2.313–2.576 | 2.317–2.563 | 2.310–2.557 | 2.307–2.568 | 2.326–2.546 |
Dy–N/Å | 2.463 | 2.464/2.468 | 2.462/2.474 | 2.456 | 2.459 |
O–Dy–O (axial)/° | 163.2 | 162.0 | 162.2 | 163.2 | 162.4 |
Dy⋯Dy/Å | 9.754 | 9.673 | 9.705 | 9.584 | 9.585 |
Slip angle θ/° | 21.91 | 22.04/23.03 | 22.44/24.49 | — | — |
dcc/Å | 3.759 | 3.835/3.713 | 3.846/3.685 | — | — |
dpp/Å | 3.428 | 3.433/3.439 | 3.418/3.442 | — | — |
dC2–C9A/Å | 3.753 | 3.832/3.707 | 3.843/3.664 | 1.658 | 1.645 |
![]() | ||
Fig. 1 The molecule structures (a, c, e and g) and 1D chain (b, d, f and h) of 1DySeCN-293, 1DySeCN-193, 1DySeCN-130 and 2DySeCN-293. |
Since one of the two 4-hpy ligands in 1DySeCN is disordered at 293 K, we measured the differential scanning calorimetry (DSC) curve to examine whether an order–disorder phase transition may occur. As shown in Fig. 2 and S3,† the DSC curve of 1DySeCN shows an exothermic peak at 276 K upon cooling and an endothermic peak at 292 K upon warming (ΔT = 16 K). Based on the endothermic peak, we can calculate the enthalpy change (ΔH) to be 1.29 kJ mol−1, and the entropy change (ΔS) to be 4.42 J mol−1 K−1. According to the Boltzmann equation ΔS = Rln(N), where R is the gas constant and N represents the number of disordered states, the value N was estimated to be 1.69, suggesting that the phase transition originates from the order–disorder transition of the axial 4-hpy ligand.34 The thermo-induced phase transition of 1DySeCN was also confirmed by variable temperature PXRD measurements (Fig. S4†). Notably, an order–disorder phase transition was not observed for the related compound DySCN down to 193 K, though one of the two axial 4-hpy ligands was also disordered in the latter.31
To investigate whether the phase transition affects the local geometry of the DyIII ion and the stacking of the anthracene groups, we determined the single-crystal structure of 1DySeCN at 193 K and named it 1DySeCN-193. Interestingly, 1DySeCN-193 crystallizes in a triclinic P space group, which is different from that measured at 293 K (Tables 1 and S1†). The asymmetric unit consists of one DyIII ion, two SeCN−, one NO3−, two depma and two 4-hpy ligands (Fig. 1c). Apparently, the phase transition upon cooling is accompanied by a shift from disorder to order in the axial 4-hpy (Fig. S5†). In addition, we also observe slight changes in bond lengths and angles. Compared with those in 1DySeCN-293, the Dy–Oax bond lengths in 1DySeCN-193 are slightly elongated while the Oax–Dy–Oax angles and Dy–Oeq bond lengths are slightly reduced (Tables 1 and S2†). The most significant structural change is probably the stacking of the anthracene units. In 1DySeCN-293, there is only one type of depma ligand in the structure, and the neighbouring anthracene rings are face-to-face π–π interacted with a dcc distance of 3.759 Å, forming an equally spaced supramolecular chain (Fig. 1b). While in 1DySeCN-193, there are two types of depma ligands (P1-An, P2-An) in the structure, each of them is face-to-face π–π interacted with its equivalent and the dcc distances are 3.713 Å for P1-An⋯P1-An and 3.835 Å for P2-An⋯P2-An, respectively, hence forming an alternatively spaced supramolecular chain (Fig. 1d).
When the measured temperature was further lowered to 130 K, the obtained 1DySeCN-130 crystallizes in the same triclinic P space group, but the cell volume decreases from 2562.4(3) Å3 at 193 K to 2536.1(5) Å3 at 130 K (Tables 1 and S1†). Again, the mononuclear molecules are stacked forming an alternatively arranged supramolecular chain (Fig. 1e and f). Compared to 1DySeCN-193, the dcc and dC2–C9A spacings of the P1-An⋯P1-An pair in 1DySeCN-130 increase slightly by 0.011 Å and 0.011 Å, while those of the P2-An⋯P2-An pair decrease by 0.028 Å and 0.043 Å, respectively (Table 1).
Clearly, lowering temperature leads to a more pronounced differentiation of the two pairs of anthracene rings and a greater deviation of each pair from the face-to-face stacking pattern. The extent to which the anthracene rings deviate from strict face-to-face stacking can be expressed in terms of the slipping angle (θ), i.e., the angle between the centroid–centroid line and the vertical line. The slipping angles of the anthracene rings are 21.910 for 1DySeCN-293, 22.040/23.030 for 1DySeCN-193, and 22.440/24.490 for 1DySeCN-130 (Fig. S6†). Indeed, the slipping angle of the anthracene pairs increases with decreasing temperature. This deviation of anthracene stacking in 1DySeCN may affect its photophysical properties.
Fig. 3b shows the PL spectra of the bulk sample of 1DySeCN as a function of irradiation time, measured on an Edinburgh FLS 980 instrument. The irradiation of the sample at 395 nm (100 mW cm−2) was performed outside the spectrometer prior to PL measurement at 365 nm. Clearly, before UV light irradiation, 1DySeCN exhibits a broad and strong emission band peaking at 550 nm (λex = 365 nm) with an average lifetime (τ) of 25.35 ns, attributed to the excimer emission of the face-to-face stacked anthracene pair (Table S5†). The quantum yield is 6.22%. After exposure to the 395 nm light, the intensity of the excimer emission decreases continuously with increasing irradiation time. At the same time, new peaks appeared at 424, 446, and 575 nm. The average lifetimes of the peaks at 422 and 446 nm are 2.3 and 4.2 ns, respectively, which are much shorter than that of the excimer emission at 550 nm and can be attributed to the π ← π* transition of the dianthracene moiety.35,36 The shoulder peak at 575 nm correspond to the f–f transition of the DyIII ion from 4F9/2 to 6H13/2, with phosphorescent lifetime of 7.01 μs (Table S5†). Clearly, photodimerization of the anthracene pairs in 1DySeSCN occurs after UV illumination, and this enhances the energy of the 3T state leading to the sensitization of the f–f transition of the dysprosium ion.
Single crystal structural analysis can provide structural information after photoinduced phase transition. By irradiating single crystals of 1DySeCN under 395 nm UV light (100 mW cm−2) at room temperature for 30 min, we obtained [DyIII(SeCN)2(NO3)(depma2)(4-hpy)2]n (2DySeCN). Although the resulting crystals were slightly cracked, they were still suitable for single crystal structure determination. Compound 2DySeCN measured at 293 K (named as 2DySeCN-293) crystallizes in the monoclinic system, space group P21/m. It has a 1D chain structure in which the equivalent DyIII ions are connected by the photo-dimerized depma2 ligands. The central C2–C9A distance in depma2 is 1.658 Å, which is remarkably shortened compared to that in 1DySeCN-293 (3.735 Å). The photodimerization of anthracene pairs also causes changes in the coordination environment of the dysprosium ion (Tables 1 and S2†). The axial Dy–O bond lengths are slightly shortened [2.208(9)/2.231(8) Å vs. 2.250(10)/2.222(11) Å in 1DySeCN-293], and the O7–Dy1–O8 angle becomes slightly smaller [162.4(6)° vs. 163.2(4)° in 1DySeCN-293]. Meanwhile, the equatorial Dy–O/N bond lengths are either elongated or shortened [2.307(7)–2.568(10) Å vs. 2.313(6)–2.576(8) Å in 1DySeCN-293]. The shortest Dy⋯Dy distance (9.584 Å) is smaller than that in 1DySeCN-293 (9.754 Å). Variations in these structural parameters explain the changes in the PL profiles of 1DySeCN-293 upon UV irradiation. Also, they will lead to changes in the magnetic properties (shown below).
Notably, the SC–SC structural transformation from 1DySeCN to 2DySeCN is reversible by annealing 2DySeCN at 105 °C. The process can be repeated at least five times with retaining the single crystallinity (Table S6 and Fig. S7†). It is worth mentioning that although one axial 4-hpy in 2DySeCN-293 is disordered, we did not see an order–disorder phase transition down to 150 K, as confirmed by the single crystal structural analysis at 193 K (Table S2 and Fig. S8†) as well as the DSC analysis in the temperature range 150–303 K (Fig. S3b†).
To ensure the completeness of the photocycloaddition reaction of the bulk sample for physical measurements, we irradiated the crystals of 1DySeCN under 395 nm UV light for 12 hours at room temperature. The purity of the resulting 2DySeCN was confirmed by its PXRD pattern which agrees well with that simulated from the single crystal data (Fig. S1†). In addition, the 1H NMR spectrum of 2DySeCN digested in DMSO-d6 solution was also measured. By integrating the area of the two sets of signals (anthracene and dianthracene), we calculated the yield of the photochemical reaction to be 95.2% (Fig. S9–S11†). The DSC curve of 2DySeCN shows an exothermic peak at 105 °C (ΔH = −51.27 kJ mol−1), indicating the occurrence of de-dimerization of the dianthracene ligand in 2DySeCN (Fig. S12†). Indeed, by annealing 2DySeCN at 105 °C for 10 min, it can be transformed back to the pristine 1DySeCN (named as 1DySeCN-re). This reversed structural transformation was supported by the PXRD, IR, UV-vis, and PL spectra measurements (Fig. 3a, S1, S13 and S14†).
The effect of temperature is more pronounced for the photochemical reaction of 1DySeCN. In general, lowering the temperature will result in a decrease in the number of molecules in the activated state, thus reducing the rate of reaction. In order to determine the lowest temperature at which the photocycloaddition reaction can occur in a reasonable amount of time, we placed 1DySeCN crystals on a hot pot and irradiated them in situ with 395 nm UV light for 2 h at different temperatures (140–200 K). The resulting samples were subjected to 1H NMR, IR and PXRD measurements (Fig. 5, S15 and S16†). As shown in Fig. 5, we calculated the yields of photochemical reactions of 1DySeCN in the temperature range of 200–140 K to be 40.8%, 21.6%, 9.5% and 6.8%, respectively. In addition, at temperatures as low as 180 K, the weak vibration peak (686 cm−1) characteristic of the dimerized anthracycline of 1DySeCN can still be observed after UV illumination (Fig. S15†). Moreover, the PXRD pattern of 1DySeCN after UV irradiation at 180 K also shows peaks corresponding to 2DySeCN (Fig. S16†). When the temperature further decreases to 160 and 140 K, these weak characteristic peaks almost disappeared in the IR and PXRD profiles.
![]() | ||
Fig. 5 The 1H NMR spectra in the range of 9.0–3.0 ppm of 1DySeCN irradiated in situ with 395 nm UV light for 2 h at different temperatures (140–200 K) in DMSO-d6. |
To examine the changes in luminescence after UV light irradiation, we selected a single crystal of 1DySeCN and studied the PL spectra after irradiation for different times at low temperatures (77–275 K) using a home-built fluorescence microscope. Due to the limited light sources available for this instrument, we used a 375 nm laser for UV irradiation and excitation with power of 9570 W cm−2. As shown in Fig. 6, the PL spectrum of the 1DySeCN single crystal at 300 K is identical to that of the bulk sample, showing a broad excimer emission peaking at 550 nm. Upon 375 nm light irradiation, photocycloaddition reactions occurred which can be monitored by its PL spectra. The intensity of the excimer emission decreased with increasing irradiation time, while the intensity of the dianthracene emission at 422 and 446 nm increased. In addition, shoulder peaks corresponding to the f–f transition of the DyIII ion appeared at 575 nm. A similar change in the PL profile was observed when the photoreaction temperature was lowered to 200–275 K. Notably, when the temperature was further reduced, the emission intensity of dianthracene and the DyIII ion became very weak at 175 K and invisible below 150 K. The results imply that at temperatures as low as 175–200 K, 1DySeCN can undergo an efficient photocycloaddition reaction concomitant with significant changes in luminescence.
Similar experiments were conducted for crystals of the dilute sample [Dy0.05Y0.95(SeCN)2(NO3)(depma)2(4-hpy)2] (1DySeCN@Y) at low temperatures (77–300 K). Upon UV light irradiation, we observed a significant reduction of the excimer emission intensity (Fig. S17†). Although the PL spectra did not show distinct characteristic emission peaks of dianthracene, the 1H NMR characterization showed the yields of photochemical reactions of 1DySeCN@Y to be 33.3% at 200 K, 25.4% at 180 K, 15.7% at 160 K and 8.3% at 140 K, respectively (Fig. S18†). Notably, the yield may be affected by the amount of sample placed on the hot stage and how it is placed.
We also tried to analyse the structure of the resulting crystal (named as 2DySeSCN-200) after irradiating 1DySeCN in situ with 395 nm UV light for 30 min at 200 K. Unfortunately, the atoms in 2DySeSCN-200 were severely disordered, so we were unable to solve the structure. However, the fact that the atoms are heavily disordered indirectly supports the occurrence of photodimerization of anthracene pairs in 1DySeCN at 200 K.
To inhibit the QTM effect, we prepared the diluted sample [Dy0.05Y0.95(SeCN)2(NO3)(depma)2(4-hpy)2] (1DySeCN@Y), and its photodimerization product [Dy0.05Y0.95(SeCN)2(NO3)(depma2) (4-hpy)2] (2DySeCN@Y). 1DySeCN@Y also exhibited reversible photochemical reactions upon UV irradiation, as confirmed by PXRD, IR, UV-vis diffuse reflectance and emission spectra as well as dc magnetic susceptibility measurements (Fig. S22† and 7a). The magnetization values are again not saturated for 1DySeCN@Y and 2DySeCN@Y at 2 K and 70 kOe (Fig. S23†). Remarkably, 1DySeCN@Y shows open butterfly-shaped hysteresis loops below 4 K characteristic of SMM behaviour (Fig. 7b and S24a†), indicating that QTM is effectively suppressed below this temperature (TB).40 The remnant and coercivity values at 2 K are 0.56 Nβ and 1200 Oe, respectively. By contrast, 2DySeCN@Y displays open hysteresis loops below 3 K with smaller remnant (0.24 Nβ) and coercivity (200 Oe) values at 2 K (Fig. 7b and S24b†). The zero-field-cooling (ZFC) and field-cooling (FC) χM vs. T curves show clear divergences at 4.2 K for 1DySeCN@Y and 3.4 K for 2DySeCN@Y, indicating that the blocking temperatures (TB) are 4.2 K and 3.4 K for the two compounds, respectively (Fig. 7c and d). Notably, the remnant, coercivity and TB values of 1DySeCN@Y are all the highest among the known photo-responsive luminescent Ln-SMMs (Table S7†).
To further understand the magnetic dynamics, we performed the alternating current (ac) susceptibility measurements of 1DySeCN and 2DySeCN under zero dc field. Both show temperature and frequency dependences of the in-phase (χ′) and out-of-phase (χ′′) susceptibility components typical for SMMs (Fig. S25 and S26†). The relaxation times (τ) were extracted using the generalized Debye model,41 and both exhibit a broad distribution of relaxation coefficients, e.g. α = 0.08–0.31 (2–22 K) for 1DySeCN and α = 0.15–0.32 (2–14 K) for 2DySeCN (Tables S8 and S9†). The lnτ vs. T−1 curves can be fitted using eqn (1), which considers the QTM, Raman, and Orbach processes.
![]() | (1) |
![]() | (2) |
The diluted samples 1DySeCN@Y and 2DySeCN@Y also show slow magnetic relaxation at zero dc field (Fig. 7e, f, S27, S28, Tables S10 and S11†). The lnτ–T−1 curves for 1DySeCN@Y and 2DySeCN@Y were fitted with eqn (2) containing Orbach and Raman processes. The resulting parameters are Ueff = 347(11) K, τ0 = 5.9(3) × 10−12 s, C = 7(1) × 10−6 K−6.23 s−1, and n = 6.2(1) for 1DySeCN@Y, and Ueff = 207(12) K, τ0 = 1.4(3) × 10−9 s, C = 1.2(4) × 10−3 K−4.56 s−1, and n = 4.6(2) for 2DySeCN@Y, respectively. Apparently, the energy barrier of the diluted sample is larger than that of the pristine samples, attributed to the suppression of the QTM effect.
The above results demonstrate that the photocycloaddition reaction of 1DySeCN and its dilute sample 1DySeCN@Y will cause remarkable changes in their SMM behavior, i.e., the reduction of effective energy barrier and the narrowing of the hysteresis loop. However, the temperature at which the SMM behaviour was observed (TB ≤ 4.2 K) differed significantly from the temperature at which the photocycloaddition reaction was observed (≥140 K). An interesting finding is that the χMT values of 1DySeCN@Y and its photodimerization product 2DySeCN@Y are quite different at about 175–200 K (Fig. 7a). This result, combined with the fact that 1DySeCN@Y can undergo photocycloaddition reaction at 175–200 K and produce notable luminescence changes (Fig. S16†), makes it possible to synergistically modulate the magnetic and luminescent properties of Ln-SMMs at the same temperature via a reversible in situ photocycloaddition reaction. Similar work has not yet been seen in the literature.
Footnotes |
† Electronic supplementary information (ESI) available: Experimental details, crystallographic details, and full characterization (SC-XRD, PXRD, IR, UV-vis and photoluminescence spectra, thermal analysis, photographs) of all described compounds. CCDC 2421525–2421527, 2424230 and 2424231. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d5sc01302j |
‡ These authors contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2025 |