Next Article in Journal
Unveiling the Multifaceted Problems Associated with Dysrhythmia
Next Article in Special Issue
Pyrimidine Schiff Bases: Synthesis, Structural Characterization and Recent Studies on Biological Activities
Previous Article in Journal
Phylogeny and Evolution of Cocconeiopsis (Cocconeidaceae) as Revealed by Complete Chloroplast and Mitochondrial Genomes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

An Effectively Uncoupled Gd8 Cluster Formed through Fixation of Atmospheric CO2 Showing Excellent Magnetocaloric Properties

1
Institute of Inorganic Chemistry (AOC), Karlsruhe Institute of Technology (KIT), Kaiserstr. 12, 76131 Karlsruhe, Germany
2
Institute of Nanotechnology (INT), Karlsruhe Institute of Technology (KIT), Kaiserstr. 12, 76131 Karlsruhe, Germany
3
Institute for Quantum Materials and Technologies (IQMT), Karlsruhe Institute of Technology (KIT), Kaiserstr. 12, 76131 Karlsruhe, Germany
4
State Key Laboratory of Rare Earth Resource Utilization, Changchun Institute of Applied Chemistry, Chinese Academy of Sciences, Renmin Street 5625, Changchun 130022, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(1), 264; https://doi.org/10.3390/ijms25010264
Submission received: 31 October 2023 / Revised: 20 December 2023 / Accepted: 21 December 2023 / Published: 23 December 2023

Abstract

:
The [Gd8(opch)8(CO3)4(H2O)8]·4H2O·10MeCN coordination cluster (1) crystallises in P 1 ¯ . The Gd8 core is held together by four bridging carbonates derived from atmospheric CO2 as well as the carboxyhydrazonyl oxygens of the 2-hydroxy-3-methoxybenzylidenepyrazine-2-carbohydrazide (H2opch) Schiff base ligands. The magnetic measurements show that the GdIII ions are effectively uncoupled as seen from the low Weiss constant of 0.05 K needed to fit the inverse susceptibility to the Curie–Weiss law. Furthermore, the magnetisation data are consistent with the Brillouin function for eight independent GdIII ions. These features lead to a magnetocaloric effect with a high efficiency which is 89% of the theoretical maximum value.

Graphical Abstract

1. Introduction

In the last few decades, molecular systems have attracted considerable attention for their possible use as magnetic refrigerants via the magnetocaloric effect (MCE) [1,2,3]. This effect, discovered by Warburg [4], describes the heating or cooling of a material through the induction of entropy changes by an applied external magnetic field. GdIII ions have been identified as optimal candidates for such clusters considering their large number of unpaired electrons and largely isotropic nature. This was shown in a Gd7 compound by Sharples et al., the first purely GdIII-containing cluster which was considered for magnetic refrigeration [3,5,6].
For the application of magnetic coolers, it is important that the field that needs to be applied in order to achieve the entropy change necessary is as low as possible [7,8]. This can be achieved when there are ferromagnetic interactions between the paramagnetic centres in the respective cluster [3].
If, however, the application of stronger fields is not an issue, another aspect becomes important, namely the efficiency of the MCE of the clusters. In order to achieve the highest possible efficiency of the MCE of a cluster, it is essential to minimise all magnetic exchange interactions between the paramagnetic centres. This leads to the entropy change observed experimentally coming close to the theoretically possible value which can be calculated using the following equation: Δ S M = n   R l n ( 2 S + 1 ) / M w , where n is the number of ions, R is the universal gas constant, S is the spin of the ion (7/2 in the case of GdIII) and Mw is the molecular weight of the cluster. This approach was recently validated by the publication of a Gd12Na6 cluster with uncoupled GdIII ions (J = −0.01 K) reaching 99% of its maximum possible entropy change [9].
In terms of efficiency, the above-mentioned equation also implies that reducing the molecular weight of the diamagnetic part of the cluster is important. Therefore, using lighter organic ligands should be preferred over inorganic ligands such as tungstate-based polyoxometalates (POMs). Whilst these can show high absolute values of entropy change, the resulting low percentages of the observed entropy change in terms of the maximum possible are a result of the heavy metal inorganic POM ligands [10].
Schiff base ligands as a type of organic ligand with ample opportunity for modification and thus the possibility to easily fine tune a system, therefore provide an excellent opportunity to obtain lanthanide clusters which could show excellent magnetocaloric properties. The ligand used in this work, 2-hydroxy-3-methoxybenzylidenepyrazine-2-carbohydrazide (H2opch) (Scheme 1), in particular has been shown to be extremely versatile. Over the last decade H2opch was used to obtain lanthanide clusters with different nuclearities, from two to eight [11,12]. The ligand was also shown to be able to accommodate both 3d and 4f ions in the same cluster [12]. This tolerance towards ions of different sizes as well as the modifiability makes H2opch the ligand of choice for the work presented here in the search for new GdIII-based magnetic refrigerants.

2. Results and Discussion

2.1. Crystallography

Here we report the synthesis, structure and magnetic properties of an octanuclear Gd8 cluster [Gd8(opch)8(CO3)4(H2O)8]·4H2O·10MeCN (1). Compound (1) crystallises in the triclinic space group P 1 ¯ with Z = 2. The core structure consists of eight GdIII ions that are arranged in a distorted square antiprism and are held together by four carbonate ligands which are presumably derived from atmospheric CO2. The eight GdIII ions are each chelated by a bridging opch2- ligand, with the negatively charged carboxyhydrazonyl oxygen of each ligand bridging to a further GdIII centre (see Scheme 1). The coordination sphere of each GdIII ion is completed by one terminal water ligand (see Figure 1). The Gd⋯Gd distances are in the range 3.9534(5)–4.0203(5) Å. All the Gd-O-Gd angles of the direct oxygen bridges range from 107.2(2) to 115.2(2)°.
The molecular structure will not be described further, as it is essentially identical to that of a Dy8 cluster previously reported by some of us [11], which crystallised in C2/c with Z = 4. By contrast, compound (1) crystallises in the lower symmetry triclinic space group P 1 ¯ (see Table S1) which leads to differences in the packing. However, the two unit cells are related. Transformation of the unit cell of (1) by the matrix (1 1 0 1 −1 0 0 0 −1) gives a C-centred cell with a = 31.412, b = 18.519, c = 28.236 Å and α = 85.18, β = 107.64, γ = 91.19°, in good agreement with the reported unit cell parameters for the Dy8 apart from the deviations of α and γ from 90°. These differences seem to be steered by the choice of solvents used in the synthetic procedures (10 MeOH and 2 H2O as lattice solvent in the reported Dy8 compound compared to 10 MeCN and 4 H2O found in the lattice of (1)).
The stability of lanthanide–carbonate compounds is exemplified by the abundance of naturally occurring lanthanide–carbonate minerals [13] and the fixation of atmospheric CO2 has been reported for a number of lanthanide-based coordination complexes [14,15,16,17,18]. It was found that basic aqueous solutions are especially prone to fix atmospheric CO2 to form bicarbonate and carbonate ions in solution [16]. In the case of the synthesis of (1), no drying procedure was used on the solvents, which therefore contain water and the Gd(NO3)3·6H2O salt was used. Additionally, N-methyldiethanolamine (mdeaH2) was added, and this acts as a base creating favourable conditions for the absorption of CO2 to form bridging carbonate ligands.

2.2. Magnetic Properties

Temperature-dependent magnetic measurements using an applied dc field of 0.1 T were performed on (1) (Figure 2). At room temperature, the χMT value of 63.80 cm3 Kmol−1 is in good agreement with the theoretical value of 63.04 cm3 Kmol−1 for eight uncoupled GdIII ions. On lowering the temperature, the value of the χMT product stays constant until ca. 25 K, below which a slight decrease was observed, reaching a minimum value of 61.96 cm3 Kmol−1 at 2 K. The extremely small decrease in χMT of only 1.84 cm3 Kmol−1 over the measured temperature range is indicative of little to no coupling present between the GdIII ions. This is further corroborated by the Curie–Weiss fit to the inverse susceptibility which results in an extremely small Weiss constant θ of 0.05 K and a Curie constant C of 63.8 cm3 Kmol−1.
This lack of interaction can be explained by the design of the ligand which has two pockets that make sure the GdIII ions are well separated. In addition, this also reduces the overall molecular weight since it precludes the need for further ligands.
Isothermal magnetisation measurements were performed at temperatures between 2 and 10 K and magnetic fields between 0 and 7 T (see Figure 3, left). At 2 K, (1) reaches its saturation magnetisation of 55.97 µB at 7 T, which is very close to the theoretical value of 56 µB for eight S = 7/2 GdIII ions. The lack of interaction between the GdIII ions is further shown by the near perfect overlap of the magnetisation at 2 K with the Brillouin curve for eight S = 7/2 ions. In ac measurements, no out-of-phase signal could be observed, which is not surprising given the complete lack of anisotropy supported by the reduced magnetisation curves (Figure 3, right).
The negligible interaction between GdIII ions suggested by the magnetic data presented above should be favourable for a strong magnetocaloric effect. Hence, the entropy change was calculated using the isothermal magnetisation measurements and Maxwell’s relation.
Δ S ( T ) Δ H =   [ M ( T ,   H ) / H ] H d H
The temperature and field dependences are shown in Figure 4. The maximum experimental entropy change of 30.36 Jkg−1 K−1 is observed at 3 K and 7 T. This equates to 89% of the maximum possible entropy change of 34.10 Jkg−1 K−1 for this compound, placing (1) amongst the most efficient Gd8 compounds for magnetic refrigeration reported in the literature [19,20,21,22,23,24].
These six Gd8 clusters all show poorer MCE behaviour with efficiencies between 73% and 81% (further details can be found in Table 1). This appears to be a result of stronger coupling between the GdIII ions as exemplified by the non-overlapping Brillouin and magnetisation curves in a [Gd8(HL)6(L)23–OH)42–OH)2(H2O)4] cluster [23]. A possible explanation for this could be the existence of direct hydroxo bridges to multiple other GdIII ions in this compound (Gd1 in this compound is directly linked to six others). In addition, in the known compounds in the literature, the GdIII ions are often interconnected by more than two oxygen-based bridges enhancing magnetic interactions. In compound (1), each GdIII ion is only connected to two others by two bridges, one carbonate and the carboxyhydrazonyl oxygen of the Schiff base ligand, which both appear to provide less favourable superexchange pathways than, for example, hydroxo and oxo bridges. This is supported by the magnetic measurements presented above and the comparison with known compounds from the literature. Furthermore, the presence of bridging carbonate has previously been identified as a means to enhance MCE behaviour [20]. Therefore, the design of the ligand to avoid strong magnetic interactions and the use of anionic co-ligands such as carbonate (irrespective of whether the carbonate forms in situ by CO2 fixation or is added as a reagent) provide a general roadmap on how to improve MCE behaviour. Schiff bases lend themselves to the task of ligand design perfectly since they can be readily modified by variation of the amine or the aldehyde used.

3. Materials and Methods

The H2opch ligand was prepared using a procedure from the literature [26]. Gd(NO3)3·6H2O and N-methyldiethanolamine (mdeaH2) were obtained from commercial sources (Sigma Aldrich) and used as delivered.
[Gd8(opch)8(CO3)4(H2O)8]·4H2O·10MeCN (1)
H2opch (0.1 mmol) was dissolved in 5 mL of MeCN and mdeaH2 (0.3 mmol) was added. Gd(NO3)3·6H2O (0.1 mmol) was separately dissolved in 5 mL of MeCN and added dropwise to the ligand/base solution. The resulting yellow solution was stirred overnight, filtered and left undisturbed for slow evaporation. After two weeks, the product was obtained as red block crystals in a yield of 17.9 mg.
IR (4000–400 cm−1): 3388 (b, s), 3054 (m), 2996 (m), 2924 (m), 2831 (m), 1603 (s), 1591 (s), 1549 (s), 1530 (s), 1520 (s), 1474 (s), 1414 (s), 1389 (s), 1337 (s), 1237 (s), 1209 (s), 1180 (m), 1154 (s), 1104 (m), 1078 (m), 1059 (w), 1027 (m), 968 (w), 919 (m), 857 (m), 844 (m), 784 (w), 769 (w), 730 (m), 641 (w), 523 (w), 481 (w), 429 (w).
C/H/N (calculated) corresponds to 10 lattice H2O: C: 32.56%, H: 2.93%, N: 11.25%
C/H/N (found): C: 32.32%, H: 2.548%, N: 11.49%.
Crystallography: A red single crystal of (1) with dimensions 0.16 × 0.14 × 0.12 mm3 was mounted on a Stoe StadiVari diffractometer equipped with a Mo GeniX 3D HF microfocus Mo-Kα source. Data were measured at 180 K, and corrected for absorption. Structure solution was by dual-space direct methods [27] and refinement by full-matrix least-squares refinement [28] against all data, on the Olex2 platform [29]. Non-H atoms were assigned anisotropic temperature factors. Geometric and rigid-bond restraints were applied to the lattice MeCN molecules as appropriate; the H-atoms of the lattice waters could not be located. Crystallographic data are summarised in Table S1.
The IR spectrum was measured on a Nicolet iS 50 with ATR attachment, the Elemental analysis was performed on a VarioELcube and magnetic SQUID measurements were performed on an MPMS-XL-7.

4. Conclusions

In conclusion we have shown the synthesis of a Gd8 cluster which forms via the fixation of atmospheric CO2 and is related to a previously reported Dy8 compound by Tang et al. [11]. As a result of the large Gd∙∙∙Gd distances, as well as the limited amount of superexchange interaction pathways compared with similar complexes in the literature, the GdIII ions are essentially uncoupled in compound (1) [19,20,21,22,23,24]. This is apparent from the magnetic susceptibility measurements. The inverse susceptibility was fitted using the Curie–Weiss law, resulting in an extremely low Weiss constant of 0.05 K. Furthermore, the magnetisation data of (1) are consistent with a Brillouin function for eight S = 7/2 ions. This, and the complete lack of anisotropy, leads to the highly efficient MCE, which is 89% of the maximum possible value. This makes compound (1) one of the most efficient refrigerants reported using larger organic ligands. This also highlights the beauty of Schiff base ligands since they can be designed to serve a given purpose. To further improve MCE behaviour, Schiff base ligands can be designed to have multiple well-separated pockets to avoid strong magnetic coupling. Furthermore, it appears that the use of anionic co-ligands such as carbonate is to be preferred over hydroxo or oxo ligands.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms25010264/s1.

Author Contributions

Conceptualization and writing, J.B., C.E.A. and A.K.P.; synthetic work and characterization, J.B. and D.S.; magnetic investigation, J.T. All authors have read and agreed to the published version of the manuscript.

Funding

The Landesgraduiertenfoerderung Baden-Wuerttemberg and the German Research Council (DFG) CRC 1573 “4f for Future”.

Data Availability Statement

Full crystallographic data and details of the structural determination have been deposited with the Cambridge Crystallographic Data Centre as supplementary publication no. CCDC 2302991. Copies of the data can be obtained, free of charge, from https://www.ccdc.cam.ac.uk/structures/.

Acknowledgments

The authors would like to thank Olaf Fuhr for his help in obtaining the crystallographic data. J.B. thanks the Landesgraduiertenfoerderung Baden-Wuerttemberg for support through a scholarship.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Tishin, A.M.; Spichkin, Y.I. The Magnetocaloric Effect and Its Application; IOP Publishing: Bristol, UK; Philadelphia, PA, USA, 2003. [Google Scholar]
  2. Evangelisti, M.; Luis, F.; de Jongh, L.J.; Affronte, M. Magnetothermal properties of molecule-based materials. J. Mater. Chem. 2006, 16, 2534–2549. [Google Scholar] [CrossRef]
  3. Zheng, Y.Z.; Zhou, G.J.; Zheng, Z.; Winpenny, R.E. Molecule-based magnetic coolers. Chem. Soc. Rev. 2014, 43, 1462–1475. [Google Scholar] [CrossRef]
  4. Warburg, E. Magnetische Untersuchungen. Ann. Phys. 1881, 249, 141–164. [Google Scholar] [CrossRef]
  5. Sharples, J.W.; Zheng, Y.Z.; Tuna, F.; McInnes, E.J.; Collison, D. Lanthanide discs chill well and relax slowly. Chem. Commun. 2011, 47, 7650–7652. [Google Scholar] [CrossRef]
  6. Sharples, J.W.; Collison, D.; McInnes, E.J.L.; Schnack, J.; Palacios, E.; Evangelisti, M. Quantum signatures of a molecular nanomagnet in direct magnetocaloric measurements. Nat. Commun. 2014, 5, 5321–5326. [Google Scholar] [CrossRef]
  7. Lorusso, G.; Sharples, J.W.; Palacios, E.; Roubeau, O.; Brechin, E.K.; Sessoli, R.; Rossin, A.; Tuna, F.; McInnes, E.J.; Collison, D.; et al. A dense metal-organic framework for enhanced magnetic refrigeration. Adv. Mater. 2013, 25, 4653–4656. [Google Scholar] [CrossRef]
  8. Mo, Z.; Gong, J.; Xie, H.; Zhang, L.; Fu, Q.; Gao, X.; Li, Z.; Shen, J. Giant low-field cryogenic magnetocaloric effect in polycrystalline LiErF4 compound. Chin. Phys. B 2023, 32, 027503. [Google Scholar] [CrossRef]
  9. Tziotzi, T.G.; Gracia, D.; Dalgarno, S.J.; Schnack, J.; Evangelisti, M.; Brechin, E.K.; Milios, C.J. A Gd12Na6 Molecular Quadruple-Wheel with a Record Magnetocaloric Effect at Low Magnetic Fields and Temperatures. J. Am. Chem. Soc. 2023, 145, 7743–7747. [Google Scholar] [CrossRef]
  10. Ibrahim, M.; Peng, Y.; Moreno-Pineda, E.; Anson, C.E.; Schnack, J.; Powell, A.K. Gd3 Triangles in a Polyoxometalate Matrix: Tuning Molecular Magnetocaloric Effects in {Gd30M8} Polyoxometalate/Cluster Hybrids Through Variation of M2+. Small Struct. 2021, 2, 2100052. [Google Scholar] [CrossRef]
  11. Tian, H.; Zhao, L.; Guo, Y.N.; Guo, Y.; Tang, J.; Liu, Z. Quadruple-CO32− bridged octanuclear dysprosium(III) compound showing single-molecule magnet behaviour. Chem. Commun. 2012, 48, 708–710. [Google Scholar] [CrossRef]
  12. Tian, H.; Ungur, L.; Zhao, L.; Ding, S.; Tang, J.; Chibotaru, L.F. Exchange Interactions Switch Tunneling: A Comparative Experimental and Theoretical Study on Relaxation Dynamics by Targeted Metal Ion Replacement. Chem. Eur. J. 2018, 24, 9928–9939. [Google Scholar] [CrossRef]
  13. Kanazawa, Y.; Kamitani, M. Rare earth minerals and resources in the world. J. Alloys Compd. 2006, 408–412, 1339–1343. [Google Scholar] [CrossRef]
  14. Natrajan, L.; Pecaut, J.; Mazzanti, M. Fixation of atmospheric CO2 by a dimeric lanthanum hydroxide complex; assembly of an unusual hexameric carbonate. Dalton Trans. 2006, 1002–1005. [Google Scholar] [CrossRef]
  15. Tang, X.-L.; Wang, W.-H.; Dou, W.; Jiang, J.; Liu, W.-S.; Qin, W.-W.; Zhang, G.-L.; Zhang, H.-R.; Yu, K.-B.; Zheng, L.-M. Olive-shaped chiral supramolecules: Simultaneous self-assembly of heptameric lanthanum clusters and carbon dioxide fixation. Angew. Chem. Int. Ed. 2009, 48, 3499–3502. [Google Scholar] [CrossRef]
  16. Langley, S.K.; Moubaraki, B.; Murray, K.S. Magnetic properties of hexanuclear lanthanide(III) clusters incorporating a central µ6-carbonate ligand derived from atmospheric CO2 fixation. Inorg. Chem. 2012, 51, 3947–3949. [Google Scholar] [CrossRef]
  17. Pineda, E.M.; Lorusso, G.; Zangana, K.H.; Palacios, E.; Schnack, J.; Evangelisti, M.; Winpenny, R.E.P.; McInnes, E.J.L. Observation of the influence of dipolar and spin frustration effects on the magnetocaloric properties of a trigonal prismatic Gd7 molecular nanomagnet. Chem. Sci. 2016, 7, 4891–4895. [Google Scholar] [CrossRef]
  18. Goura, J.; Colacio, E.; Herrera, J.M.; Suturina, E.A.; Kuprov, I.; Lan, Y.; Wernsdorfer, W.; Chandrasekhar, V. Heterometallic Zn3Ln3 Ensembles Containing (µ6-CO3) Ligand and Triangular Disposition of Ln3+ ions: Analysis of Single-Molecule Toroic (SMT) and Single-Molecule Magnet (SMM) Behavior. Chem. Eur. J. 2017, 23, 16621–16636. [Google Scholar] [CrossRef]
  19. Bala, S.; Adhikary, A.; Bhattacharya, S.; Bishwas, M.S.; Poddar, P.; Mondal, R. Ln8 (Ln = Gd, Ho, Er, Yb) Butterfly Core-Exhibiting Magnetocaloric Effect and Field-Induced SMM Behavior for Er Analogue. ChemistrySelect 2017, 2, 11341–11345. [Google Scholar] [CrossRef]
  20. Wu, J.; Li, X.L.; Zhao, L.; Guo, M.; Tang, J. Enhancement of Magnetocaloric Effect through Fixation of Carbon Dioxide: Molecular Assembly from Ln4 to Ln4 Cluster Pairs. Inorg. Chem. 2017, 56, 4104–4111. [Google Scholar] [CrossRef] [PubMed]
  21. Cui, C.; Ju, W.; Luo, X.; Lin, Q.; Cao, J.; Xu, Y. A Series of Lanthanide Compounds Constructed from Ln8 Rings Exhibiting Large Magnetocaloric Effect and Interesting Luminescence. Inorg. Chem. 2018, 57, 8608–8614. [Google Scholar] [CrossRef] [PubMed]
  22. Li, L.-F.; Kuang, W.-W.; Li, Y.-M.; Zhu, L.-L.; Xu, Y.; Yang, P.-P. A series of new octanuclear Ln8 clusters: Magnetic studies reveal a significant cryogenic magnetocaloric effect and slow magnetic relaxation. New J. Chem. 2019, 43, 1617–1625. [Google Scholar] [CrossRef]
  23. Kalita, P.; Goura, J.; Nayak, P.; Colacio, E.; Chandrasekhar, V. Octanuclear {Ln8} complexes: Magneto-caloric effect in the {Gd8} analogue. J. Chem. Sci. 2021, 133, 82. [Google Scholar] [CrossRef]
  24. Li, J.N.; Li, N.F.; Wang, J.L.; Liu, X.M.; Ping, Q.D.; Zang, T.T.; Mei, H.; Xu, Y. A new family of boat-shaped Ln8 clusters exhibiting the magnetocaloric effect and slow magnetic relaxation. Dalton Trans. 2021, 50, 13925–13931. [Google Scholar] [CrossRef] [PubMed]
  25. Chen, Y.-C.; Qin, L.; Meng, Z.-S.; Yang, D.-F.; Wu, C.; Fu, Z.; Zheng, Y.-Z.; Liu, J.-L.; Tarasenko, R.; Orendac, M.; et al. Study of a magnetic-cooling material Gd(OH)CO3. J. Mater. Chem. A 2014, 2, 9851–9858. [Google Scholar] [CrossRef]
  26. Vergara, F.M.; Lima, C.H.; Henriques, M.; Candea, A.L.; Lourenco, M.C.; de, L. Ferreira, M.; Kaiser, C.R.; de Souza, M.V. Synthesis and antimycobacterial activity of N’-[(E)-(monosubstituted-benzylidene)]-2-pyrazinecarbohydrazide derivatives. Eur. J. Med. Chem. 2009, 44, 4954–4959. [Google Scholar] [CrossRef] [PubMed]
  27. Sheldrick, G.M. SHELXT—Integrated space-group and crystal-structure determination. Acta Cryst. A 2015, 71, 3–8. [Google Scholar] [CrossRef]
  28. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Cryst. C 2015, 71, 3–8. [Google Scholar] [CrossRef]
  29. Dolomanov, O.V.; Bourhis, L.J.; Gildea, R.J.; Howard, J.A.K.; Puschmann, H. OLEX2: A complete structure solution, refinement and analysis program. J. Appl. Cryst. 2009, 42, 339–341. [Google Scholar] [CrossRef]
Scheme 1. The neutral H2opch Schiff base ligand (left) and the coordination modes of the (opch)2− ligands in (1) (right).
Scheme 1. The neutral H2opch Schiff base ligand (left) and the coordination modes of the (opch)2− ligands in (1) (right).
Ijms 25 00264 sch001
Figure 1. Two views of the molecular structure (top) and the core structure of (1) (bottom) highlighting the bridging modes of the carbonate ligands with orange bonds.
Figure 1. Two views of the molecular structure (top) and the core structure of (1) (bottom) highlighting the bridging modes of the carbonate ligands with orange bonds.
Ijms 25 00264 g001
Figure 2. The χMT vs. T plot (black dots), the inverse susceptibility vs. T plot (red circles) and the Curie–Weiss fit (blue line) for (1) between 300 and 2 K.
Figure 2. The χMT vs. T plot (black dots), the inverse susceptibility vs. T plot (red circles) and the Curie–Weiss fit (blue line) for (1) between 300 and 2 K.
Ijms 25 00264 g002
Figure 3. (a) Magnetisation and (b) reduced magnetisation curves of (1) at temperatures between 2 and 10 K and 0 and 7 T.
Figure 3. (a) Magnetisation and (b) reduced magnetisation curves of (1) at temperatures between 2 and 10 K and 0 and 7 T.
Ijms 25 00264 g003
Figure 4. Molar entropy change calculated from isothermal magnetisation measurements vs. T (a) and B (b).
Figure 4. Molar entropy change calculated from isothermal magnetisation measurements vs. T (a) and B (b).
Ijms 25 00264 g004
Table 1. Formulae, molecular weights and entropy changes with the efficiency in brackets, as well as the magnetic fields and temperatures the entropy change was measured at for different Gd8 compounds in the literature using larger organic ligands.
Table 1. Formulae, molecular weights and entropy changes with the efficiency in brackets, as well as the magnetic fields and temperatures the entropy change was measured at for different Gd8 compounds in the literature using larger organic ligands.
CompoundMW (g/mol)−ΔSM (J/kgK)B (T)T (K)Reference
[Gd8(opch)8(CO3)4(H2O)8]·4H2O·10MeCN4286.7630.36 (89%)73this work
[Gd83-OH)4(L1)4(DEA)4Cl4](DMF)2(MeOH)3404.0031.4 (77%)93[19]
[Gd83-OH)4(CO3)2L4(PhCOO)8]3982.6028.38 (78%)72[20]
[Gd8(CH2OHCH2OH)8(SO4)12]·2(C2H7N)·2H2O3012.3036.86 (80%)72[21]
[Gd83-O)4(L)8(CH3COO)4(CO3)2]·15H2O3463.8532.49 (81%)72[22]
[Gd8(HL)6(L)23-OH)4(µ2-OH)2(H2O)4]3777.8825.5 (73%)53[23]
[Gd8(IN)143-OH)82-OH)2(H2O)8]·11H2O3479.8131.77 (80%)72[24]
Note: These values can be compared with complexes using smaller organic or inorganic ligands, such as [Gd12Na6(OAc)25(HCO2)5(CO3)6(H2O)12]·9H2O·0.5MeCN which is highly efficient (29.3 J/kgK (99%)) [9], and extended solids, such as [Gd(OH)CO3] which reaches extremely high entropy changes (66.4 J/kgK) [25].
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Braun, J.; Seufert, D.; Anson, C.E.; Tang, J.; Powell, A.K. An Effectively Uncoupled Gd8 Cluster Formed through Fixation of Atmospheric CO2 Showing Excellent Magnetocaloric Properties. Int. J. Mol. Sci. 2024, 25, 264. https://doi.org/10.3390/ijms25010264

AMA Style

Braun J, Seufert D, Anson CE, Tang J, Powell AK. An Effectively Uncoupled Gd8 Cluster Formed through Fixation of Atmospheric CO2 Showing Excellent Magnetocaloric Properties. International Journal of Molecular Sciences. 2024; 25(1):264. https://doi.org/10.3390/ijms25010264

Chicago/Turabian Style

Braun, Jonas, Daniel Seufert, Christopher E. Anson, Jinkui Tang, and Annie K. Powell. 2024. "An Effectively Uncoupled Gd8 Cluster Formed through Fixation of Atmospheric CO2 Showing Excellent Magnetocaloric Properties" International Journal of Molecular Sciences 25, no. 1: 264. https://doi.org/10.3390/ijms25010264

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop