Next Article in Journal
Chloride-Induced Stress Corrosion Cracking of Friction Stir-Welded 304L Stainless Steel: Effect of Microstructure and Temperature
Previous Article in Journal
Superlattice Symmetries Reveal Electronic Topological Transition in CaC6 with Pressure
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis, Structure and Magnetic Properties of Low-Dimensional Copper(II) trans-1,4-cyclohexanedicarboxylate

by
Pavel A. Demakov
1,*,
Anna A. Ovchinnikova
1,2,
Pavel V. Dorovatovskii
3,
Vladimir A. Lazarenko
3,
Alexander N. Lavrov
1,
Danil N. Dybtsev
1 and
Vladimir P. Fedin
1,*
1
Nikolaev Institute of Inorganic Chemistry, Siberian Branch of the Russian Academy of Sciences, 630090 Novosibirsk, Russia
2
Department of Natural Sciences, Novosibirsk State University, 630090 Novosibirsk, Russia
3
National Research Centre “Kurchatov Institute”, 123182 Moscow, Russia
*
Authors to whom correspondence should be addressed.
Crystals 2024, 14(6), 555; https://doi.org/10.3390/cryst14060555
Submission received: 23 May 2024 / Revised: 11 June 2024 / Accepted: 13 June 2024 / Published: 15 June 2024
(This article belongs to the Section Hybrid and Composite Crystalline Materials)

Abstract

:
A reaction between copper(II) nitrate and trans-1,4-cyclohexanedicarboxylic acid (H2chdc) carried out under hydrothermal conditions led to a new metal-organic coordination polymer [Cu2(Hchdc)2(chdc)]n. According to single-crystal XRD data, the compound is based on bi-nuclear paddlewheel-type carboxylate blocks that are joined with polymeric chains due to the (μ312) coordination of carboxylate groups. The chains are interconnected by chdc2− bridging ligands into layers containing free COOH groups of terminal Hchdc. The neighboring layers adopt a RCOOH···OOCR hydrogen bond-assisted arrangement into a dense-packed structure. Magnetization measurements showed the presence of a strong antiferromagnetic exchange interaction (J/kB = −495 K) inside the bi-nuclear blocks. At the same time, no significant interaction was found between the {-Cu2(OOCR)4-} units in spite of their polymeric in-chain packing. Patterns of magnetic behavior of [Cu2(Hchdc)2(chdc)]n were thoroughly analyzed and explained from a structural point of view.

Graphical Abstract

1. Introduction

Metal-organic coordination polymers (MOCPs) have attracted significant research attention due to their broad structural design [1,2,3] and targeted modulation of properties that is possible by varying functional metal centers [4], organic ligands [5], and guest substrates [6]. Polynuclear and polymeric blocks consisting of unpaired electron-rich metal ions can present valuable magnetic properties (such as ferro- [7], antiferro- [8], and ferrimagnets [9], as well as slow magnetic relaxation [10,11]), complex luminescence patterns [12,13], redox behavior [14], and unique chemical bonding features [15]. In particular, such metal centers are of interest for the synthesis of MOCPs with designable and unique properties and unveil the usage of coordination polymers in diverse magnetic materials [16], in highly efficient luminophores and sensors [4,17], for data storage [18], and for production of multifunctional and “smart” materials [19].
Copper(II) is one of the readily available ions with valuable functional characteristics, and it is widely used in MOCP chemistry [20,21]. The redox activity and optical properties of Cu2+ are applicable in the design of catalysts and photocatalysts [22,23], while the relative structural rigidity of its coordination environment increases the predictability of the polymer lattice [24]. The paramagnetism of Cu2+ ion (S = ½) can provide deep insights into the structural dynamics and the nature of adsorption centers in mechanistic studies [25,26,27], by means of physical methods, and allows us to achieve an extremely strong magnetic exchange in its polymeric complexes [28,29]. In this work, a new MOCP with the formula [Cu2(Hchdc)2(chdc)]n (I), containing differently charged anions of trans-1,4-cyclohexanedicarboxylic acid (H2chdc), being also readily available but mildly spread in coordination chemistry [30], as ligands, was structurally characterized. A (μ312) coordination of the carboxylates and hydrogen bonds between them and free carboxylic groups induce a formation of a corrugated layered structure based on 1D chains of magnetoactive Cu2+ ions. However, magnetic measurements demonstrated that these copper-carboxylate chains represent a set of weakly interacting antiferromagnetic dimers with a very strong (J/kB = −495 K) inner exchange and a pronounced influence of structural defects regarding their low-temperature magnetization behavior.

2. Materials and Methods

2.1. Materials

Copper(II) nitrate trihydrate (analytical reagent; Chimmed, Moscow, Russian Federation), trans-1,4-cyclohexanedicarboxylic acid (H2chdc, >97.0%; TCI, Tokyo, Japan), and acetone (reagent grade; Vekton, Saint Petersburg, Russian Federation) were used as received. Distilled water was used in all synthetic experiments.

2.2. Instruments

Bruker Scimitar FTS 2000 spectrometer (Billerica, MA, USA) was used to obtain infrared (IR) spectra in the range of 4000−400 cm−1 with KBr pellets. VarioMICROcube analyzer (Elementar Analysensysteme GmbH, Hanau, Germany) was used for elemental CHNS analyses. Bruker D8 Advance diffractometer (Cu-Kα radiation, λ = 1.54178 Å; Billerica, MA, USA) was used for powder X-ray diffraction (PXRD) data acquisition at room temperature. Netzsch TG 209 F1 Iris device (Selb, Germany) was used for thermogravimetric analysis under Ar flow (30 cm3∙min−1) at a 10 K∙min−1 heating rate.
A Quantum Design MPMS-XL (San Diego, CA, USA) SQUID magnetometer was used to measure the magnetic susceptibility of [Cu2(Hchdc)2(chdc)] (I) in the temperature range of 1.77–300 K under applied magnetic fields H = 0–10 kOe. The measured values of the total molar susceptibility, χ = M/H, were corrected for the temperature-independent diamagnetic core contribution, χd, calculated using the Pascal’s additive scheme, which allowed us to determine the paramagnetic component of the magnetic susceptibility, χp(T). To check for the presence of ferromagnetic trace impurities in the sample, we measured isothermal M(H) dependencies at several temperatures and took M(T) data in various magnetic fields; a weak FM contribution, χFM(T), was indeed detected in this way and subtracted from the data. Due to the low magnetic susceptibility of the sample and its low amount available, 4 cycles of measurements were carried out in the range of 1.77–300 K at both heating and cooling to increase the accuracy of measurements and confirm their reproducibility.

2.3. Synthetic Methods

Synthesis of [Cu2(Hchdc)2(chdc)]n (I). Cu(NO3)2·3H2O (1.67 g, 6.9 mmol), H2chdc (1.46 g, 8.5 mmol), and 25.0 mL of water were mixed in a Teflon liner (50 mL general volume). The liner was placed into a stainless steel autoclave and heated at 120 °C during 48 h with 50 °C·h−1 heating and cooling rates. After cooling, the obtained white-blue crystalline precipitate was washed with water (~20 mL) and then with acetone (~5 mL) and dried in air. Yield: 5%. IR spectrum main bands (KBr, cm−1): 3191 ν(O–HCOOH), 2944 ν(C(sp3)–HCHring), 2853 ν(C(sp3)–HCH2ring), 1728 ν(COCOOH), 1585 νas(COO), and 1416 νs(COO). Elemental CHN analysis, calculated for [Cu2(Hchdc)2(chdc)] (C24H32O12Cu2) (%): C 45.1; H 5.0; N 0.0. Found (%): C 45.0; H 5.0; N 0.0.

2.4. Single Crystal X-ray Diffraction Details

Diffraction data for single crystals of I were collected on the ‘Belok’ beamline [31,32] (λ = 0.74503 Å) of the National Research Center ‘Kurchatov Institute’ (Moscow, Russian Federation) using a Rayonix SX165 CCD detector. XDS program (Version 10 January 2022) package [33] was used for data indexing, integration, scaling, and absorption correction. Dual-space algorithm (SHELXT [34]) was used for structure solution, and the full-matrix least squares technique (SHELXL [35]) was used for structure refinement. Anisotropic approximation was applied for all atoms, except hydrogens. Positions of C-bonded hydrogen atoms were calculated geometrically and refined in the riding model. Positions of O-bonded H atoms were found directly and further refined with the application of DFIX restraints for stable refinement. A non-merohedral twinning with an orientation matrix (−1 0 0 0 −1 0 −0.426 −0.199 1) and second component weight (BASF) of ca. 0.34 was found and analyzed using PLATON [36] software (Version 13 May 2024) after finding a structural model and its primary refinement. Details for single-crystal structure determination experiments and structure refinements are summarized in Table 1. CCDC 2309813 (I) entry contains the supplementary crystallographic data for this paper. These data can be obtained free of charge from The Cambridge Crystallographic Data Center at https://www.ccdc.cam.ac.uk/structures/ (accessed on 12 June 2024).

3. Results and Discussion

3.1. Synthesis

Compound [Cu2(Hchdc)2(chdc)]n (I) was obtained in a single-crystalline phase in low yield by heating a mixture of copper(II) nitrate and H2chdc in water in a stainless steel autoclave with a Teflon liner at 120 °C for 48 h. After cooling the reaction mixture to room temperature, the solid mixture contained blue single crystals of I and colorless crystals of unreacted trans-1,4-cyclohexanedicarboxylic acid. Then, H2chdc was removed by washing the precipitate with acetone to obtain I in a pure form. Attempts to increase the target product yield by adding potassium hydroxide or several weak organic bases (such as 1,4-diazabicyclo[2.2.2]octane, urotropine, triethylamine, urea, N,N-dimethylformamide) to the reaction mixture, as well as by increasing the temperature of the synthesis, led to the admixtures of the previously reported [Cu2(chdc)2] [37] or unidentified copper-containing products.

3.2. Crystal Structure

According to SCXRD results, I crystallized in the triclinic system with the P¯1 space group. I contains two independent Cu(II) ions. Both Cu(1) and Cu(2) have a similar square/pyramidal coordination polyhedron, consisting of five oxygen atoms from five COO groups, two of which have a (μ211) coordination, and three more adopt the (μ312) coordination (Figure 1a). Cu–O bond lengths are very close for both independent metal ions and lie in the range of 1.947(7)–1.996(7) Å for O atoms of pyramid bases and are 2.204(6) Å, 2.210(6) Å for the pyramid vertices. Such coordination modes of carboxylate groups lead to the formation of paddlewheel blocks condensed through the O atoms of μ3-COO groups into polymeric {-Cu2(OOCR)4-}n chains lying along a crystallographic axis. The Cu…Cu distances inside paddlewheels are 2.571(2) Å and 2.576(2) Å, and the shortest interblock Cu…Cu distances are 3.21–3.22 Å, which can suggest strong exchange interactions between paramagnetic metal ions both inside the bi-nuclear blocks and between them—along polymeric chains. Nearby chains are interlinked along the b axis by bridging trans-1,4-cyclohexanedicarboxylates (chdc2−) into layers (Figure 1b). Monoprotonated Hchdc ligands containing uncoordinated COOH groups are coordinated to the chains almost perpendicularly to the ab planes (Figure 1c), forming interlayer H(COOH)···O(COO) hydrogen bonds with the corresponding oxygen–oxygen distances of 2.72–2.73 Å. The resulting three-dimensional packing of layers in I does not contain any significant voids.

3.3. Phase Purity and Characterization

The phase purity of I was confirmed by powder X-ray diffraction (PXRD, Figure 2a). Elemental (CHN) analysis results are in full agreement with the chemical formula [Cu2(Hchdc)2(chdc)]n obtained from the SCXRD data. The IR spectrum of I contains characteristic absorption bands of C(sp3)–H stretching vibrations at 2944 cm−1 and 2853 cm−1, C=O stretching vibrations in the free COOH group at 1728 cm−1, symmetric and antisymmetric vibrations of the coordinated carboxyl group at 1416 cm−1 and 1585 cm−1, respectively, and a broad absorption band of O–H stretching vibrations at ~3200 cm−1. According to thermogravimetric analysis (TGA) data, the sample is stable up to 300 °C. The first step of weight loss is observed at 345 °C (Figure 2b). The residual weight after the first decomposition step is 48%, corresponding well to the presumable formula Cu2O(chdc) for the intermediate decomposition product (49.0% theoretical weight). The second weight loss step is observed at 390 °C. The final weight residue after the second stage is 22% and corresponds well to Cu2O as the final product for the decomposition of I (22.4% theoretical weight) in an inert atmosphere. Such a stepwise shape of the TGA curve is similar to the examples within the literature for other “acidic” trans-1,4-cyclohexanedicarboxylates, such as non-isostructural [Mn2(Hchdc)2(chdc)]n [38].

3.4. Magnetization Data

The magnetic susceptibility χ = M/H of complex I, measured during thermal cycling of the sample in the temperature range of 1.77–300 K, did not depend on either its thermomagnetic history or the magnetic field strength, except for at lower temperatures, where the field dependence χ(H) fitted well to the Brillouin function for paramagnetic ions with spin S = 1/2. The paramagnetic component of the magnetic susceptibility χp (Figure 3a), obtained after subtracting the temperature-independent diamagnetic contribution, was ≈1.5·10−3 emu/mol at 300 K and gradually decreased almost to zero upon cooling the sample from room temperature to 50–60 K. Such a pronounced drop in magnetic susceptibility indicates not only the presence of strong antiferromagnetic (AF) interactions between Cu2+ ions but also the formation of a gap in the spectrum of spin excitations in the system; without such a gap, significant magnetic susceptibility would persist, even in a completely ordered AF state. The occurrence of a spin gap along with the smooth decrease in χp are known to be the typical manifestations of AF-coupled dimers [39] and dimer-derived structures. This observation appears in perfect agreement with the crystal structure of complex I based on polymeric chains of bi-nuclear Cu(II)-carboxylate blocks (Figure 1). The strength of the AF interaction can be seen clearly in the χpT(T) graph (Figure 3b), where χpT is almost zero at temperatures below 60–70 K and reaches only half the value that would be observed for the two Cu2+ ions in [Cu2(Hchdc)2(chdc)]n at room temperature—in the absence of an interaction between them.
In the low-temperature region (T < 60 K), an increase in χp is observed (Figure 3a), which is associated with the presence of impurities retaining their paramagnetic state down to the minimal temperature T = 1.77 K. An impurity paramagnetism of sample I apparently arises from the decomposition of bi-nuclear blocks in near-surface crystallite defects with the formation of single Cu2+ ions [40]. As can be seen from Figure 3a, there is virtually no overlap of temperature ranges at points where significant contributions to the magnetic susceptibility of AF dimers (T ≥ 70 K) and paramagnetic impurities (T ≤ 50 K) are observed. This allows the concentration and magnetic parameters of single Cu2+ ions to be determined by simply approximating the low-temperature χp(T) data (Figure 4) using the Curie–Weiss dependence χ CW ( T ) = N A μ eff 2 / 3 k B ( T θ ) —where NA and kB are the Avogadro number and the Boltzmann constant, respectively, μeff is the effective magnetic moment of a single Cu2+ ion, and θ is the Weiss constant—reflecting its interaction with neighboring copper ions. In turn, the analysis of the χp(T) dependence in the region T ≥ 70 K has shown good agreement with the behavior theoretically predicted for AF Cu2+–Cu2+ dimers [39]. Accordingly, a quantitative analysis of the data over the entire temperature range was carried out using the modified Bleaney–Bowers formula [39], which takes into account the presence of AF dimers and monomer molecules:
χp(T) = (1 − δ)χdim + 2δχCW = (1 − δ)(2NAμB2g2/3kBT)/[1 + exp(–J/kBT)/3] + 2δχCW,
where χdim and χCW are the magnetic susceptibilities of dimers and monomers, respectively; δ is the mole fraction of dimers transformed into monomers; g is the g-factor of Cu2+ ions (assumed to be the same for dimers and monomers); and J is the exchange interaction energy between copper ions in dimers described by the Hamiltonian H = J S 1 S 2 . As can be seen in Figure 5, the experimental data can be approximated very well using Formula (1) with parameters J/kB ≈ −495 K; g ≈ 2.19; δ ≈ 0.45 mol. %.
Since the dimer blocks are assembled into polymer chains within the crystal structure of complex I, a more rigorous analysis of the magnetic properties should be carried out using the model of partially dimerized chains described by the Hamiltonian H = J i n / 2 ( S 2 i 1 S 2 i + α S 2 i S 2 i + 1 ) , where the inter-dimer interaction J’ = αJ is also considered in addition to the intra-dimer Cu2+–Cu2+ exchange J [41]. However, upon applying this model, no further improvement exceeding the experimental error was obtained in the description of the χp(T) data, which indicates that the interaction between the blocks is very weak (J’ << J).
Despite the apparent consistency between magnetization and structural data, one may wonder whether the identification of bi-nuclear paddlewheel fragments of polymer chains as AF Cu2+-Cu2+ dimers, weakly coupled with each other magnetically, is actually correct. Indeed, at a first glance, the exchange-interaction pathways through Cu-O-Cu bonds (inset to Figure 5) may seem preferable over the Cu-O-C-O-Cu ones. However, this is not the case; both the strong AF coupling in the paddlewheel blocks and the weak superexchange interaction through Cu-O-Cu bonds of the geometry in complex I are confirmed by many sources of data within the literature on similar compounds. The most clear example is the compound [Cu2Y2L10(H2O)4·3H2O]n, where HL = trans-2-butenoic (crotonic) acid, in which {Cu2(OOCR)4} paddlewheel units in polymer chains are spatially separated by diamagnetic bi-nuclear {Y2(OOCR)6} blocks and are thus magnetically isolated from each other [42]. The value of the AF exchange interaction in isolated copper(II)-based paddlewheel dimer units was found to be J/kB ≈ −486 K [41], appearing very close to the value −495 K obtained in this work. Even closer magnetic parameters (J/kB ≈ −493 K; g ≈ 2.188) were defined for isolated {Cu2(OOCR)4} blocks in ref. [37]. In turn, studies of di-nuclear Cu(II) complexes with Cu2O2 cores possessing a geometry similar to complex I have revealed only a weak (|J/kB| ~ 5 K) exchange interaction through Cu-O-Cu bonds, whose sign is dependent on the Cu-O-Cu angle [43].
Among the compounds based on {-Cu2(OOCR)4-}n polymeric chains [28,37,44], the closest ones to complex I parameters (J/kB = −485 K; g = 2.207) were reported for layered compound [Cu2(cis,cis-1,3,5-Hchtc)2]n [37], where 1,3,5-H3chtc = 1,3,5-cyclohexanetricarboxylic acid. For all the referenced chain-based compounds, the insignificance of the exchange interaction between bi-nuclear blocks was emphasized [28,37,44], and only the authors in ref. [28] qualitatively determined the value of J’/kB ≈ 8.5 K, which appears to be almost two orders of magnitude smaller than the interaction within the dimer units (J/kB). It should be noted that in all the listed works, impurity paramagnetism (described by the Curie–Weiss model) was observed at T < 50 K, and it was typically subtracted empirically and not analyzed in detail. We can suggest that such a feature of complex I is associated with the destruction of paddlewheel units in the near-surface defects of crystallites, forming monomer paramagnetic ions. According to the parameter δ ≈ 0.45 mol. % (reflecting the fraction of decomposed dimers as ≈1/220), the average number of paddlewheel blocks in a single {-Cu2(OOCR)4-}n chain should be n ≈ 220. Using the cell parameter a = 5.16 Å, whose crystallographic vector is aligned parallel to the chains, the average length of defect-free copper(II)-carboxylate chains in complex I can be estimated as (220·0.516 ≈ 114) nm, which is quite large and consistent with high crystallinity shown by PXRD results.

4. Conclusions

To summarize, a new metal-organic coordination polymer of copper(II), which contains differently charged trans-1,4-cyclohexanedicarboxylate anions, was synthesized. The compound is based on two-dimensional coordination layers, while its magnetically active substructure is represented by one-dimensional copper-carboxylate chains. The packing of layers within the crystal structure is supported by hydrogen bonds between protonated (uncoordinated) carboxyl groups and bound carboxylate anions. Magnetization measurements have shown the presence of strong antiferromagnetic interactions (J/kB = −495 K) in {Cu2(OOCR)4} paddlewheel units without any significant exchange between even the closest bi-nuclear fragments.

Author Contributions

Conceptualization, P.A.D. and V.P.F.; methodology, P.A.D., A.N.L. and D.N.D.; software, A.N.L., P.V.D. and V.A.L.; validation, P.A.D. and V.P.F.; formal analysis, A.N.L.; investigation, A.A.O. (synthesis, characterization, graphing), A.N.L. (magnetization measurements, graphing), P.V.D. (single-crystal XRD), V.A.L. (single-crystal XRD); resources, D.N.D. and V.P.F.; data curation, P.V.D., V.A.L. and V.P.F.; writing—original draft preparation, P.A.D., A.A.O. and A.N.L.; writing—review and editing, D.N.D. and V.P.F.; visualization, P.A.D., A.A.O. and A.N.L.; supervision, P.A.D.; project administration, V.P.F.; funding acquisition, V.P.F. All authors have read and agreed to the published version of the manuscript.

Funding

The authors are grateful to the Ministry of Science and Higher Education of the Russian Federation for financial support (Agreement No. 075-15-2022-263), and for providing access to the large-scale research facility “EXAFS spectroscopy beamline”.

Data Availability Statement

CCDC 2309813 entry contains the supplementary crystallographic data for this paper. This original data presented in the study are openly available in The Cambridge Crystallographic Data Center at https://www.ccdc.cam.ac.uk/structures/ (accessed on 12 June 2024). Other characterization method original datasets will be made available on request from the corresponding authors due to privacy.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Yaghi, O.M.; O’Keeffe, M.; Ockwig, N.W.; Chae, H.K.; Eddaoudi, M.; Kim, J. Reticular synthesis and the design of new materials. Nature 2003, 423, 705–714. [Google Scholar] [CrossRef]
  2. Barsukova, M.; Sapianik, A.; Guillerm, V.; Shkurenko, A.; Shaikh, A.C.; Parvatkar, P.; Bhatt, P.M.; Bonneau, M.; Alhaji, A.; Shekhah, O.; et al. Face-directed assembly of tailored isoreticular MOFs using centring structure-directing agents. Nat. Synth. 2024, 3, 33–46. [Google Scholar] [CrossRef]
  3. Zhang, Q.; Yan, S.; Yan, X.; Lv, Y. Recent advances in metal-organic frameworks: Synthesis, application and toxicity. Sci. Total Environ. 2023, 902, 165944. [Google Scholar] [CrossRef]
  4. Lunev, A.M.; Belousov, Y.V. Luminescent sensor materials based on rare-earth element complexes for detecting cations, anions, and small molecules. Russ. Chem. Bull. 2022, 71, 825–857. [Google Scholar] [CrossRef]
  5. Zaguzin, A.S.; Sukhikh, T.; Sokolov, M.N.; Fedin, V.P.; Adonin, S.A. Zn(II) Three-Dimensional Metal-Organic Frameworks Based on 2,5-Diiodoterephthalate and N,N Linkers: Structures and Features of Sorption Behavior. Inorganics 2023, 11, 192. [Google Scholar] [CrossRef]
  6. Sun, Z.; Khurshid, A.; Sohail, M.; Qiu, W.; Cao, D.; Su, S.J. Encapsulation of Dyes in Luminescent Metal-Organic Frameworks for White Light Emitting Diodes. Nanomaterials 2021, 11, 2761. [Google Scholar] [CrossRef]
  7. Wu, Y.; Cheng, S.; Liu, J.; Yang, G.; Wang, Y.Y. New porous Co(II)-based metal-organic framework including 1D ferromagnetic chains with highly selective gas adsorption and slow magnetic relaxation. J. Solid State Chem. 2019, 276, 226–231. [Google Scholar] [CrossRef]
  8. Gong, T.; Li, P.; Sui, Q.; Zhou, L.J.; Yang, N.N.; Gao, E.Q. Switchable Ferro-, Ferri-, and Antiferromagnetic States in a Piezo- and Hydrochromic Metal–Organic Framework. Inorg. Chem. 2018, 57, 6791–6794. [Google Scholar] [CrossRef] [PubMed]
  9. Mahata, P.; Sen, D.; Natarajan, S. A three-dimensional metal–organic framework with a distorted Kagome related layer showing canted antiferromagnetic behaviour. Chem. Commun. 2008, 2008, 1278–1280. [Google Scholar] [CrossRef]
  10. Kuppusamy, S.K.; Moreno-Pineda, E.; Nonat, A.M.; Heinrich, B.; Karmazin, L.; Charbonniere, L.J.; Ruben, M. Binuclear Lanthanide Complexes Based on 4-Picoline-N-oxide: From Sensitized Luminescence to Single-Molecule Magnet Characteristics. Cryst. Growth Des. 2023, 23, 1084–1094. [Google Scholar] [CrossRef]
  11. Matyukhina, A.K.; Zorina-Tikhonova, E.N.; Goloveshkin, A.S.; Babeshkin, K.A.; Efimov, N.N.; Kiskin, M.A.; Eremenko, I.L. Field-Induced Slow Magnetic Relaxation in CoII Cyclopropane-1,1-dicarboxylates. Molecules 2022, 27, 6537. [Google Scholar] [CrossRef] [PubMed]
  12. Salaam, J.; Tabti, L.; Bahamyirou, S.; Lecointre, A.; Hernandez Alba, O.; Jeannin, O.; Camerel, F.; Cianférani, S.; Bentouhami, E.; Nonat, A.M.; et al. Formation of Mono- and Polynuclear Luminescent Lanthanide Complexes based on the Coordination of Preorganized Phosphonated Pyridines. Inorg. Chem. 2018, 57, 6095–6106. [Google Scholar] [CrossRef] [PubMed]
  13. Ivanova, E.A.; Smirnova, K.S.; Pozdnyakov, I.P.; Potapov, A.S.; Lider, E.V. Photoluminescent Lanthanide(III) Coordination Polymers with Bis(1,2,4-Triazol-1-yl)Methane Linker. Inorganics 2023, 11, 317. [Google Scholar] [CrossRef]
  14. Litvinova, Y.M.; Gayfulin, Y.M.; Kovalenko, K.A.; Samsonenko, D.G.; Van Leusen, J.; Korolkov, I.V.; Fedin, V.P.; Mironov, Y.V. Multifunctional Metal–Organic Frameworks Based on Redox-Active Rhenium Octahedral Clusters. Inorg. Chem. 2018, 57, 2072–2084. [Google Scholar] [CrossRef] [PubMed]
  15. Afonin, M.Y.; Konokhova, A.Y.; Dmitriev, A.A.; Abramov, P.A.; Kuratieva, N.V.; Sukhikh, T.S.; Kompankov, N.B.; Gritsan, N.P.; Konchenko, S.N. Chromium–Lanthanide Complexes Containing the Cr═P═Cr Fragment: Synthesis, Characterization, and Computational Study. Inorg. Chem. 2023, 62, 10110–10119. [Google Scholar] [CrossRef] [PubMed]
  16. Mínguez Espallargas, G.; Coronado, E. Magnetic functionalities in MOFs: From the framework to the pore. Chem. Soc. Rev. 2018, 47, 533–557. [Google Scholar] [CrossRef] [PubMed]
  17. Yu, X.; Ryadun, A.A.; Pavlov, D.I.; Guselnikova, T.Y.; Potapov, A.S.; Fedin, V.P. Highly Luminescent Lanthanide Metal-Organic Frameworks with Tunable Color for Nanomolar Detection of Iron(III), Ofloxacin and Gossypol and Anti-counterfeiting Applications. Angew. Chem. Int. Ed. 2023, 135, e202306680. [Google Scholar] [CrossRef]
  18. Yang, C.; Dong, R.; Wang, M.; Petkov, P.S.; Zhang, Z.; Wang, M.; Han, P.; Ballabio, M.; Bräuninger, S.A.; Liao, Z.; et al. A semiconducting layered metal-organic framework magnet. Nat. Commun. 2019, 10, 3260. [Google Scholar] [CrossRef] [PubMed]
  19. Zhang, T.; Hu, Y.Q.; Mo, Z.W.; Liao, P.Q.; Sakiyama, H.; Han, T.; Chen, X.M.; Zheng, Y.Z. Cobalt(II) Magnetic Metal–Organic Framework with an Effective Kagomé Lattice, Large Surface Area, and High Spin-Canted Ordering Temperature. ACS Appl. Mater. Interfaces 2017, 9, 38181–38186. [Google Scholar] [CrossRef]
  20. Cun, J.E.; Fan, X.; Pan, Q.; Gao, W.; Luo, K.; He, B.; Pu, Y. Copper-based metal–organic frameworks for biomedical applications. Adv. Colloid Interface Sci. 2022, 305, 102686. [Google Scholar] [CrossRef]
  21. Stirk, A.J.; Wilson, B.H.; O’Keefe, C.A.; Amarne, H.; Zhu, K.; Schurko, R.W.; Loeb, S.J. Applying reticular synthesis to the design of Cu-based MOFs with mechanically interlocked linkers. Nano Res. 2021, 14, 417–422. [Google Scholar] [CrossRef]
  22. An, B.; Li, Z.; Song, Y.; Zhang, J.; Zeng, L.; Wang, C.; Lin, W. Cooperative copper centres in a metal–organic framework for selective conversion of CO2 to ethanol. Nat. Catal. 2019, 2, 709–717. [Google Scholar] [CrossRef]
  23. Nwosu, U.; Siahrostami, S. Copper-based metal–organic frameworks for CO2 reduction: Selectivity trends, design paradigms, and perspectives. Catal. Sci. Technol. 2023, 13, 3740–3761. [Google Scholar] [CrossRef]
  24. Liu, J.; Lukose, B.; Shekhah, O.; Arslan, H.K.; Weidler, P.; Gliemann, H.; Bräse, S.; Grosjean, S.; Godt, A.; Feng, X.; et al. A novel series of isoreticular metal organic frameworks: Realizing metastable structures by liquid phase epitaxy. Sci. Rep. 2012, 2, 921. [Google Scholar] [CrossRef] [PubMed]
  25. Demakov, P.A.; Poryvaev, A.S.; Kovalenko, K.A.; Samsonenko, D.G.; Fedin, M.V.; Fedin, V.P.; Dybtsev, D.N. Structural Dynamics and Adsorption Properties of the Breathing Microporous Aliphatic Metal–Organic Framework. Inorg. Chem. 2020, 59, 15724–15732. [Google Scholar] [CrossRef] [PubMed]
  26. Mendt, M.; Ehrling, S.; Senkovska, I.; Kaskel, S.; Pöppl, A. Synthesis and Characterization of Cu–Ni Mixed Metal Paddlewheels Occurring in the Metal–Organic Framework DUT-8(Ni0.98Cu0.02) for Monitoring Open-Closed-Pore Phase Transitions by X-Band Continuous Wave Electron Paramagnetic Resonance Spectroscopy. Inorg. Chem. 2019, 58, 4561–4573. [Google Scholar] [CrossRef] [PubMed]
  27. Kolganov, A.A.; Gabrienko, A.A.; Stepanov, A.G. Reaction of methane with benzene and CO on Cu-modified ZSM-5 zeolite investigated by 13C MAS NMR spectroscopy. Chem. Phys. Lett. 2023, 810, 140188. [Google Scholar] [CrossRef]
  28. Perec, M.; Baggio, R.; Sartoris, R.P.; Santana, R.C.; Peña, O.; Calvo, R. Magnetism and Structure in Chains of Copper Dinuclear Paddlewheel Units. Inorg. Chem. 2010, 49, 695–703. [Google Scholar] [CrossRef] [PubMed]
  29. Julve, M.; Gleizes, A.; Chamoreau, L.M.; Ruiz, E.; Verdaguer, M. Antiferromagnetic Interactions in Copper(II) µ-Oxalato Dinuclear Complexes: The Role of the Counterion. Eur. J. Inorg. Chem. 2018, 2018, 509–516. [Google Scholar] [CrossRef]
  30. Demakov, P.A. Properties of Aliphatic Ligand-Based Metal–Organic Frameworks. Polymers 2023, 15, 2891. [Google Scholar] [CrossRef]
  31. Svetogorov, R.D.; Dorovatovskii, P.V.; Lazarenko, V.A. Belok/XSA Diffraction Beamline for Studying Crystalline Samples at Kurchatov Synchrotron Radiation Source. Cryst. Res. Technol. 2020, 55, 1900184. [Google Scholar] [CrossRef]
  32. Lazarenko, V.A.; Dorovatovskii, P.V.; Zubavichus, Y.V.; Burlov, A.S.; Koshchienko, Y.V.; Vlasenko, V.G.; Khrustalev, V.N. High-Throughput Small-Molecule Crystallography at the ‘Belok’ Beamline of the Kurchatov Synchrotron Radiation Source: Transition Metal Complexes with Azomethine Ligands as a Case Study. Crystals 2017, 7, 325. [Google Scholar] [CrossRef]
  33. Kabsch, W. XDS. Acta Cryst. 2010, D66, 125–132. [Google Scholar] [CrossRef] [PubMed]
  34. Sheldrick, G.M. SHELXT–Integrated space-group and crystal-structure determination. Acta Cryst. 2015, A71, 3–8. [Google Scholar] [CrossRef] [PubMed]
  35. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Cryst. 2015, C71, 3–8. [Google Scholar] [CrossRef] [PubMed]
  36. Spek, A.L. Single-crystal structure validation with the program PLATON. J. Appl. Crystallogr. 2003, 36, 7–13. [Google Scholar] [CrossRef]
  37. Kumagai, H.; Akita-Tanaka, M.; Inoue, K.; Takahashi, K.; Kobayashi, H.; Vilminot, S.; Kurmoo, M. Metal−Organic Frameworks from Copper Dimers with cis- and trans-1,4-Cyclohexanedicarboxylate and cis,cis-1,3,5-Cyclohexanetricarboxylate. Inorg. Chem. 2007, 46, 5949–5956. [Google Scholar] [CrossRef] [PubMed]
  38. Demakov, P.A.; Bogomyakov, A.S.; Urlukov, A.S.; Andreeva, A.Y.; Samsonenko, D.G.; Dybtsev, D.N.; Fedin, V.P. Transition Metal Coordination Polymers with Trans-1,4-Cyclohexanedicarboxylate: Acidity-Controlled Synthesis, Structures and Properties. Materials 2020, 13, 486. [Google Scholar] [CrossRef] [PubMed]
  39. Bleaney, B.; Bowers, K.D. Anomalous paramagnetism of copper acetate. Proc. R. Soc. Lond. 1952, A214, 451–465. [Google Scholar] [CrossRef]
  40. Shen, L.; Yang, S.W.; Xiang, S.; Liu, T.; Zhao, B.; Ng, M.F.; Göettlicher, J.; Yi, J.; Li, S.; Wang, L.; et al. Origin of Long-Range Ferromagnetic Ordering in Metal–Organic Frameworks with Antiferromagnetic Dimeric-Cu(II) Building Units. J. Am. Chem. Soc. 2012, 134, 17286–17290. [Google Scholar] [CrossRef]
  41. Hatfield, W.E. New magnetic and structural results for uniformly spaced, alternatingly spaced, and ladder-like copper (II) linear chain compounds. J. Appl. Phys. 1981, 52, 1985–1990. [Google Scholar] [CrossRef]
  42. Calvo, R.; Rapp, R.E.; Chagas, E.; Sartoris, R.P.; Baggio, R.; Garland, M.T.; Perec, M. 1-D Polymers with Alternate Cu2 and Ln2 Units (Ln = Gd, Er, Y) and Carboxylate Linkages. Inorg. Chem. 2008, 47, 10389–10397. [Google Scholar] [CrossRef] [PubMed]
  43. Baggio, R.; Calvo, R.; Garland, M.T.; Peña, O.; Perec, M.; Slep, L.D. A new copper (II) di-μ2-carboxylato bridged dinuclear complex: [Cu(oda)phen]2·6H2O (oda = oxydiacetate, phen = phenanthroline). Inorg. Chem. Commun. 2007, 10, 1249–1252. [Google Scholar] [CrossRef]
  44. Agterberg, F.P.; Provó Kluit, H.A.; Driessen, W.L.; Oevering, H.; Buijs, W.; Lakin, M.T.; Spek, A.L.; Reedijk, J. Dinuclear Paddle-Wheel Copper(II) Carboxylates in the Catalytic Oxidation of Carboxylic Acids. Unusual Polymeric Chains Found in the Single-Crystal X-ray Structures of [Tetrakis(μ-1-phenylcyclopropane-1-carboxylato-O,O‘)bis(ethanol-O)dicopper(II)] and catena-Poly[[bis(μ-diphenylacetato-O:O‘)dicopper](μ3-diphenylacetato-1-O:2-O‘:1‘-O‘)(μ3-diphenylacetato-1-O:2-O‘:2‘-O‘)]. Inorg. Chem. 1997, 36, 4321–4328. [Google Scholar] [CrossRef] [PubMed]
Figure 1. {Cu2(OOCR)4} bi-nuclear blocks and their connection in the structure of I (a). Coordination network in I; view along c axis (b). Three-dimensional package of I; view along a axis (c). Copper atoms are shown in blue; oxygen atoms are shown in red. Hydrogen bonds are shown by orange dashed lines.
Figure 1. {Cu2(OOCR)4} bi-nuclear blocks and their connection in the structure of I (a). Coordination network in I; view along c axis (b). Three-dimensional package of I; view along a axis (c). Copper atoms are shown in blue; oxygen atoms are shown in red. Hydrogen bonds are shown by orange dashed lines.
Crystals 14 00555 g001aCrystals 14 00555 g001b
Figure 2. Powder X-ray diffraction pattern of sample I compared to the theoretical one (a). TGA curve and dm/dT plot for I (b).
Figure 2. Powder X-ray diffraction pattern of sample I compared to the theoretical one (a). TGA curve and dm/dT plot for I (b).
Crystals 14 00555 g002
Figure 3. Temperature dependences of χp (a) and χpT (b) of complex I measured in a magnetic field H = 10 kOe. Each curve was obtained by averaging two series of measurements. The dashed lines in graph (b) show the levels corresponding to one ion with spin S = 1/2 and g-factors g = 2.0; 2.19.
Figure 3. Temperature dependences of χp (a) and χpT (b) of complex I measured in a magnetic field H = 10 kOe. Each curve was obtained by averaging two series of measurements. The dashed lines in graph (b) show the levels corresponding to one ion with spin S = 1/2 and g-factors g = 2.0; 2.19.
Crystals 14 00555 g003
Figure 4. Temperature dependences of 1/χp of complex I measured in a magnetic field H = 10 kOe. The dashed line shows the approximation of low-temperature data using the Curie–Weiss dependence.
Figure 4. Temperature dependences of 1/χp of complex I measured in a magnetic field H = 10 kOe. The dashed line shows the approximation of low-temperature data using the Curie–Weiss dependence.
Crystals 14 00555 g004
Figure 5. Experimental χp(T) data for complex I (symbols) and their approximation by the theoretical expression obtained using the Bleaney–Bowers model (solid line). Contributions from dimers and monomers are shown by the dashed and dash-dotted lines, respectively. The inset illustrates the chain built from AF dimers with strong intra-dimer coupling J, and a weak interaction between dimers J’.
Figure 5. Experimental χp(T) data for complex I (symbols) and their approximation by the theoretical expression obtained using the Bleaney–Bowers model (solid line). Contributions from dimers and monomers are shown by the dashed and dash-dotted lines, respectively. The inset illustrates the chain built from AF dimers with strong intra-dimer coupling J, and a weak interaction between dimers J’.
Crystals 14 00555 g005
Table 1. Single-crystal X-ray diffraction and structure refinement details.
Table 1. Single-crystal X-ray diffraction and structure refinement details.
ParameterI
Chemical formulaC12H16CuO6
Mr, g·mol−1319.79
Crystal systemTriclinic
Space groupP¯1
a, Å5.160(1)
b, Å10.820(2)
c, Å22.630(5)
Temperature, K100
α, °86.95(3)
β, °86.90(3)
γ, °83.34(3)
V, Å31251.7(4)
Z4
Dcalc, g·cm−31.697
μ, mm−11.99
F(000)660
Crystal size, mm0.04 × 0.02 × 0.02
θ range for data collection, °2.0 ≤ θ ≤ 26.55
Index ranges−6 ≤ h ≤ 6; −13 ≤ k ≤ 13; −3 ≤ l ≤ 27
No. of reflections: measured/independent/
observed [I > 2σ(I)]
4376/4376/4145
Rint
Goodness-of-fit on F21.221
Final R indices [I > 2σ(I)]R1 = 0.1024; wR2 = 0.2961
Final R indices [all data]R1 = 0.1045; wR2 = 0.2970
Largest diff. peak, hole, e·Å−31.90, −1.75
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

Demakov, P.A.; Ovchinnikova, A.A.; Dorovatovskii, P.V.; Lazarenko, V.A.; Lavrov, A.N.; Dybtsev, D.N.; Fedin, V.P. Synthesis, Structure and Magnetic Properties of Low-Dimensional Copper(II) trans-1,4-cyclohexanedicarboxylate. Crystals 2024, 14, 555. https://doi.org/10.3390/cryst14060555

AMA Style

Demakov PA, Ovchinnikova AA, Dorovatovskii PV, Lazarenko VA, Lavrov AN, Dybtsev DN, Fedin VP. Synthesis, Structure and Magnetic Properties of Low-Dimensional Copper(II) trans-1,4-cyclohexanedicarboxylate. Crystals. 2024; 14(6):555. https://doi.org/10.3390/cryst14060555

Chicago/Turabian Style

Demakov, Pavel A., Anna A. Ovchinnikova, Pavel V. Dorovatovskii, Vladimir A. Lazarenko, Alexander N. Lavrov, Danil N. Dybtsev, and Vladimir P. Fedin. 2024. "Synthesis, Structure and Magnetic Properties of Low-Dimensional Copper(II) trans-1,4-cyclohexanedicarboxylate" Crystals 14, no. 6: 555. https://doi.org/10.3390/cryst14060555

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