Skip to main content
ACS AuthorChoice logoLink to ACS AuthorChoice
. 2023 Aug 7;62(33):13253–13276. doi: 10.1021/acs.inorgchem.3c01506

Syntheses and Characterization of Main Group, Transition Metal, Lanthanide, and Actinide Complexes of Bidentate Acylpyrazolone Ligands

Thomas Mies †,*, Andrew J P White , Henry S Rzepa , Luciano Barluzzi , Mohit Devgan , Richard A Layfield , Anthony G M Barrett
PMCID: PMC10445273  PMID: 37549423

Abstract

graphic file with name ic3c01506_0030.jpg

The synthesis of acylpyrazolone salts and their complexes of main group elements, transition metals, lanthanides, and actinides are described and characterized inter alia by means of single-crystal X-ray crystallography, NMR, and IR spectroscopies. The complexes consist of two, three, or four acylprazolone ligands bound to the metal atom, resulting in a structurally diverse set of coordination complexes with (distorted) octahedral, pentagonal-bipyramidal, or antiprismatic arrangements. Several complexes proved to be polymeric in the solid state including heterobimetallic sodium/lanthanide coordination polymers. A selection of the polymeric compounds was analyzed via TG/DTA measurements to establish their stability. The ligands, in turn, were readily synthesized in good yields from commercially available hydrazine hydrochloride salts. These findings demonstrate that acylpyrazolone ligands can form complexes with metals of varying ionic radii, highlighted by their utility in other areas such as analytical and metal organic framework chemistry.

Short abstract

The syntheses and structural analyses of the representative main group, transition metal, lanthanide, and actinide pyrazolonato complexes highlighted in red are presented.

Introduction

Alternating oligo-carbonyl compounds, such as β-diketones, are well understood bi-, tri-, and higher order-ligands for metal ion complexation due to their O,O–bidentate chelating coordination properties.1 The acetylacetone (acac)H ligand (1), Scheme 1, is the archetypical β-diketone, and its longstanding history in metal organic chemistry is mirrored by the multitude of utilization ranging from fundamental research in coordination chemistry to applications, including catalysis, metal extraction, analytical chemistry, water purification, and material science.14

Scheme 1. Structures of acac (1) and a Generalized Acyl Pyrazolone (2) Including the Various Tautomeric Forms AD of the Derived Anion.

Scheme 1

Fusing the β-dicarbonyl system within a heterocycle, such as pyrazole, gives rise to a subclass of β-diketones, which are known as acyl-pyrazolones (2), Scheme 1.1b,5 Upon deprotonation, these molecules benefit from an additional potential coordination site on the nitrogen atom (A), next to the β-diketone-derived forms of the anion BD, and are found to have generally lower pKa-values than the parent β-diketones. Consequently, ligands derived from acyl-pyrazolones (2) were shown to be excellent O,O-chelators in metal ion extractions via complexes with high-stability constants and low solubility in aqueous media.5,6 Furthermore, pyrazolones are commonly used scaffolds in medicinal chemistry, and multiple studies on the biological evaluation of metal-pyrazolone complexes as antitumor agents have been reported.7,8

Previously, we reported the synthesis and characterization by single-crystal X-ray and micro-ED structure determinations of a calcium-pyrazolonato complex.9 Herein, we wish to report more comprehensive studies on the synthesis and characterization of several acyl-pyrazolones and their coordination chemistry with the representative main group, transition metals, lanthanides and uranium salts.

Results and Discussion

Syntheses of Acyl-Pyrazolone Ligands

The ligand precursors were prepared from commercially available hydrazine derivatives as described by Holzer and Taylor.10 Reaction of hydrazine hydrate or phenylhydrazine with diethyl ethoxymethylenemalonate (4) gave the unprotected acylpyrazolone (5) and phenyl-protected acylpyrazolone (6) in 58 and 92% yields, respectively (Scheme 2). Reaction of 4-methoxybenzyl-(PMB-)hydrazine hydrochloride (7) with malonate 4 in water in the presence of potassium carbonate gave acylpyrazolone (8) in 93% yield. Saponification with 2 M aqueous potassium hydroxide, followed by acid mediated decarboxylation gave pyrazolone (9) (72%). Subsequent reaction with benzoyl chloride and calcium hydroxide in 1,4-dioxane gave benzoyl-pyrazolone (10) (68%).

Scheme 2. Syntheses of Acyl-Pyrazolone Ligands 5, 6, 8, and 10.

Scheme 2

Syntheses of Acyl-Pyrazolone Metal Complexes

Previous syntheses of the acyl-pyrazolone main group and transition metal complexes involved the in situ preparation of the alkaline earth metal complexes by reaction with alkali metal hydroxides or methoxides (Li, Na, K), followed by salt metathesis with a second metal salt to provide the target complexes.5a,5b In the context of nucleoside syntheses by N-glycosylation reactions, we had occasion to further study these complexation reactions and we initially examined the synthesis of alkaline earth metal complexes (M = Mg, Ca). Since these studies provided us with several interesting metal complexes, we extended our studies to other main group metals, and to transition metals, as well as lanthanide and uranium salts. In the interests of concise presentation, the results are presented with summaries of syntheses and characterizations and with detailed discussions of solid-state structures in the following section. A structurally diverse set of complexes were obtained from the reaction of pyrazolones 5, 6, 8, and 10 with the following metal precursors: sodium methoxide, magnesium chloride, calcium trifluoromethanesulfonate, scandium trifluoromethanesulfonate, yttrium chloride, titanium tetramethoxide, zirconium tetrachloride, zirconocene dichloride, dirhodium tetraacetate, manganous acetate, ferric chloride hexahydrate, zinc acetate dihydrate, cupric acetate hydrate, nickel(II) chloride hexahydrate, lanthanum trifluoromethanesulfonate, dysprosium trifluoromethanesulfonate, ytterbium trifluoromethanesulfonate, and uranyl nitrate hexahydrate (Scheme 3 and Table 1).

Scheme 3. Overview of the General Method to Prepare Acylpyrazolonato Complexes with Di-, Tri-, and Tetravalent Metals Listed in Table 1.

Scheme 3

Table 1. Syntheses of Metal Pyrazolonato Complexes from Pyrazolones 5, 6, 8, and 10a.

graphic file with name ic3c01506_0028.jpg

graphic file with name ic3c01506_0029.jpg

a

M, L, X, R1, and R2 are defined in the Table. Reaction stoichiometries (pyrazolone:metal precursor) were as follows: 1:1 (entries 1–4); 2:1 (entries 5–12, 19–22, 27, 28, 31–34, 39–50, and 63–66); 3:1 (entries 13–18, 35–38, and 51–62); and 4:1 (entries 23–26, 29, and 30). L = MeOH, EtOH, pyridine, or H2O (entries 1–4, 51, 52, 54–56, 58–60, and 62–66).

The 1H-NMR spectra of each sodium complex showed coordination-induced shifts for the imine-proton signals. For example, with phenyl-pyrazolone (6), the signal shifted from δ = 7.81 to 7.65 ppm in CD3OD upon coordination with the sodium cation (12). Similar shifts (δ = 7.75 to 7.48 ppm, 7.77 to 7.38 ppm, and 7.83 to 7.46 ppm CD3OD) were observed for pyrazolones 11, 13, and 14, respectively, upon formation of the sodium complexes. Related shifts in the 13C{1H}-NMR signal of the pyrazolone carbonyl group were also observed. DSC measurements of the sodium pyrazolone complexes 1114 revealed phase transitions between 67 and 72 °C, consistent with initial desolvation and possible alteration from an α-phase to a β-phase of the complexes. This phase transition required enthalpies of 3.98 J g–1 (complex 11), 4.62 J g–1 (complex 12), 2.80 J g–1 (complex 13), and 3.53 J g–1 (complex 14).

The reactions of the proligands with alkaline-earth metal salts were conducted according to method A (Scheme 3). Coordination-induced shifts were once more observed in the 1H- and 13C{1H}-NMR spectra, while in the solid state, the compounds are predominantly isostructural. The Mg(II) and Ca(II) ions occupy octahedral coordination sites and reside on a center of symmetry with two pyrazolonato ligands in the xy-plane, and with two solvent molecules (EtOH or H2O/MeOH) in the axial positions (see the following discussions of X-ray crystal structures).

The scandium and yttrium complexes were prepared according to method B (Scheme 3). Yttrium-complex 27 was found to be isomorphous with the corresponding Ph-protected pyrazolone lanthanum complex 62 and ytterbium-complex 70, c.f. Table 1, underscoring the structural similarities of these metals. Consequently, the yttrium-complex will be discussed in comparison to the lanthanide complexes.

The syntheses of titanium(IV) complexes were executed according to method A (2 equiv of pyrazolone with titanium tetramethoxide). The 1H-NMR spectra indicated characteristic coordination-induced shifts for the imine protons. Furthermore, for titanium complexes 29 and 31, the benzylic protons are split into a doublet of doublets, indicative of coordination to the metal. The analogous zirconium(IV) complexes were prepared according to method C (Scheme 3). Furthermore, two zirconium half-sandwich complexes were prepared from Cp2ZrCl2, which was allowed to react with ligands 8 and 6 in the presence of pyridine, giving complexes 37 and 38 in 57% and 32%, respectively.

The reactions of dirhodium tetraacetate with the pyrazalones 8 and 6 at elevated temperatures in acetonitrile resulted in coordination of the pyrazalones ligands at the axial positions giving complexes 39 and 40 in yields of 98% and 93%, respectively. In neither reaction was any substitution of the bridging acetate ligands observed.

Previous studies reported on acylpyrazolone-Fe(III) and -Mn(II) complexes by Okafor revealed high-spin-configured complexes in an octahedral environment as evidenced by magnetic susceptibility studies. Using an Evans/Johnson-Matthey balance, the acylpyrazolone-Fe(III) complexes showed μeff values of between 5.70 and 5.95 BM,11a,11b while the acylpyrazolone-Mn(II) complexes had μeff values of between 5.30 and 5.60 BM.11c To further study these observations, reactions of the d5 metals manganese(II) and iron(III) were examined with pyrazolones 5, 6, 8, and 10. The manganese complexes 4144 were synthesized from manganese(II) diacetate according to method A, while the iron complexes 4548 were prepared from ferric chloride hexahydrate according to method B.

The copper(II) and nickel(II) complexes 5356 and 5760, prepared according to method A, are expected to contain at least one axial solvent molecule. NMR spectroscopic characterization was uninformative since only broad peaks were observed, consistent with the paramagnetic nature of the complexes. Various experiments in common NMR solvents (such as CDCl3, CD3OD, CD3CN, and (CD3)2CO) failed to provide a spectrum of the complexes and only when dissolved in (CD3)2SO did the spectrum consist of broad peaks, consistent with the paramagnetic arrangement of the d8 electrons. When recrystallized from pyridine, nickel complex 60 was obtained with an octahedral coordination with two pyridine molecules in the axial position replacing ligated ethanol (see the following discussions of X-ray crystal structures). The remaining structures are tentatively assigned as containing two axial ethanol solvent molecules; however, the bulk sample material might be of a polymeric nature,9 or without ligated solvent molecules. X-Ray crystal structures of the Cu- or Zn-complexes 49, 54, and 55 were found to be polymeric in nature. Ni-pyrazolonato complexes have previously been reported with axial DMF, MeOH, or EtOH ligands.12

The reactions of the pyrazolone proligands 5, 6, 8, and 10 with a representative selection of f-block elements were examined, and complexes were prepared according to method B. While complexes 61, 62, and 64 gave reasonably well-resolved NMR-spectra, the benzoyl-substituted complex (63) showed a broadening of peaks. It is assumed that this could originate from geometrical constraints of the complex, leading to restricted and/or slow bond rotation of the substituents.13

DSC measurements of the metal organic framework (MOF)-complex 61 revealed phase transitions at 68 °C, consistent with initial desolvation of the MOF, which required a phase transition enthalpy of 1.43 J g–1. Cross validation with TG/DTA measurements confirmed an initial mass-change of 4% (desolvation) followed by negligible changes on the subsequent heating cycles (0.4 and 0.2%). DSC measurements of MOF-complex 65 revealed phase transitions at 68 °C, consistent with initial desolvation of the MOF which required a phase-transition enthalpy of 1.66 J g–1. Cross validation with TG/DTA measurements confirmed an initial mass-change of 1% (desolvation) followed by negligible changes on the subsequent heating cycles (0.16 and 0.15%).

Discussion of Solid-State Structures

Alkali-Metal Complexes

Selected bond lengths and angles of sodium complexes 11 (Figure 1), and 12 and 14 (Figure 2) are given in Table 2. The C–O bond lengths of the pyrazolonato-ligand were generally comparable to literature values (1.256(2) and 1.245(2) Å, respectively), and thus as reported by Bochkarev, longer than the C=O double bond (1.21 Å).14 Similar bond lengths for Na–O bonds (2.335(1) and 2.352(1) Å) were reported for a related dimethoxyethane-coordinated sodium-pyrazolonato complex.14 The dihedral angles of the O–Na–O coordination per pyrazolone are slightly larger than the structure reported by Bochkarev and co-workers (77.40(4)°).14

Figure 1.

Figure 1

Molecular structure of Na-complex 11.

Figure 2.

Figure 2

Crystal structures of Na-complexes 12 and 14.

Table 2. Selected Bond Lengths and Angles for the Alkali-Metal Complexes.
selected bond lengths [Å] and angles [°] 11 selected bond lengths [Å] and angles [°] 12 selected bond lengths [Å] and angles [°] 14
M–O3/O3A 2.3480(9) M–O3/O3A 2.353(13) M–O3 2.3133(17)
M–O15/O15A 2.4555(10) M–O12/O12A 2.377(13)/2.444(13) M–O6 2.3542(19)
M–O20/O20A 2.4496(12) M–O20/O20A 2.383(13)/2.443(13) M–O16 2.5258(19)
C–O3/O3A 1.2522(16) C–O3/O3A 1.250(18) M–N1/N12 2.375(2)/2.399(2)
C–O15/O15A 1.2227(16) C–O12/O12A 1.24(2) C–O3 1.282(3)
O3–M–O15 80.33(3) O3–M–O12 79.0(4) C–O6/16 1.228(3)/1.217(3)
O3–M–O15A 90.67(3) O3–M–O12A 81.9(4) O3–M–O6 83.21(6)
O3–M–O20 85.74(4) O3–M–O20 85.3(4) O3–M–O16 95.02(7)
O3–M–O20A 94.26(4) O3–M–O20A 114.5(5) O3–M–O12x 100.62(7)

The sodium complex 11 has a polymeric structure as a result of intermolecular H-bonding. The repeating unit has an octahedral coordination of the sodium ion with two pyrazolone ligands and two aquo ligands. One of the pyrazolone ligands is charge bearing, while the other one is neutral, and a proton was located in the difference map on the terminal nitrogen. This proton is H-bonding to an adjacent pyrazolone ligand via an N–H–N bond of 2.627(2) Å, giving rise to an H-bonded polymer. Additional X-ray structure determinations of other sodium complexes were measured following crystallization from methanol or ethanol. In the solid state, these complexes also formed polymeric structures, with the sodium center being octahedrally coordinated. The phenyl-protected pyrazolone-sodium complex 12 formed polymers via bridging of the acylpyrazolone ligand with an adjacent repeating unit. The structure resulted in Na···Na separations of 3.110(2) Å. The unprotected-pyrazolone sodium complex 14, on the other hand, formed polymers by binding to both oxygens of the β-diketone moiety, as well as one of the nitrogens of adjacent pyrazolone rings. This bonding of Na+ to the nitrogen atoms was nonspecific, and coordinative bonds via both the N1 or N2 position were found. The Na···Na separation in complex 12 is 3.6155(18) Å, comparable with those of complex 14 (3.110(2) Å) and literature data (3.573(1) Å).14

Alkaline-Earth Metal Complexes

Selected bond lengths and angles for the magnesium and calcium complexes are summarized in Table 3. The metal ion in each complex occupies a slightly distorted octahedral geometry (Figure 3). The bond lengths and angles of the magnesium(II)-complexes compare well to the ones found in [Mg(acac)2(H2O)2] (acac–Mg: 2.040 Å, 2.027 Å; Mg–OH2: 2.148 Å)15 with the noted distortion along the equatorial Mg–O bond and shorter Mg–O(H)Et bond. Notably, in the benzoyl-pyrazolonato magnesium complex 17, the Mg—O(=C) and the Mg–O(H)Et bonds are equidistant (2.107(2) and 2.099(2) Å, respectively).

Table 3. Selected Bond Lengths and Angles for the Octahedral Alkaline-Earth Complexes.
selected bond lengths [Å] and angles [°] 15 19 16 17 21 18 22
M–O1/O1A 2.0211(10) 2.3116(11) 2.0473(19) 2.050(2) 2.2895(16) 2.0635(9) 2.3048
M–O2/O2A 2.1295(11) 2.3724(11) 2.075(3) 2.107(2) 2.3370(17) 2.0866(9) 2.3055
M–O3/O3A 2.0835(11) 2.3296(12) 2.107(3) 2.099(2) 2.3490(18) 2.1063(10) 2.3623
C–O1/O1A 1.2725(18) 1.2669(19) 1.268(4) 1.268(4) 1.261(3) 1.2816(16) 1.2761
C–O2/O2A 1.2376(18) 1.2385(19) 1.233(4) 1.255(4) 1.253(3) 1.2386(16) 1.2330
O1–M–O2 89.77(4) 81.28(4) 89.49(9) 89.34(8) 80.68(6) 88.92(3) 81.06
O1–M–O2A 90.23(4) 98.72(4) 90.51(9) 90.66(8) 99.32(6) 91.08(3) 98.94
O1–M–O3 91.32(4) 87.28(4) 89.83(10) 92.40(8) 90.99(6) 90.15(4) 89.87
O1–M–O3A 88.68(4) 92.72(4) 90.17(10) 87.60(8) 89.01(6) 89.85(4) 90.13
Figure 3.

Figure 3

General structure and atom labeling in Mg- and Ca-complexes 15–19, 21, and 22 listed in Table 3.

The isomorphous calcium(II)-complexes have comparable bond lengths and angles relative to reported literature data of 2.29–2.32 Å, which is typical for Ca–O pyrazolonato bond lengths8f,16 and also compares well to reported bond lengths for the Ca–solvent ligand bonds (Ca–O(H)Et bond: 2.35–2.38 Å).8f,16 The found bond lengths also compare well to the ones found in [Ca(acac)2(H2O)2] (acac–Ca: 2.336(2) Å, 2.320(2) Å; Ca–OH2: 2.356(2) Å).17 The phenyl-pyrazolone complexes on the other hand gave rise to different structures, while the magnesium complex 16 possessed the expected octahedral coordination sphere (data in Table 2), the corresponding calcium complex 20 had a dimeric, pentagonal-bipyramidal coordination around the calcium center, with two pyrazolonato ligands, a water and ethanol solvent molecule, Figure 4.

Figure 4.

Figure 4

Molecular structure of calcium(II)-complex 20. Selected bond lengths: Ca1–O3: 2.4047(11) Å; Ca1–O12: 2.4314(11) Å; Ca1–O23: 2.4179(10) Å (part of the Ca2O2-ring); Ca1–O32: 2.4397(11) Å; Ca1–O40: 2.4553(13) Å; Ca1–O50: 2.3314(11) Å. The Ca2O2-ring was completed by bond Ca1–O23A: 2.3829(10) Å, giving a Ca1–Ca1A separation of 3.7523(6) Å. Selected bond angles in the Ca2O2-ring are: O23A–Ca1–O23: 77.19(4)°; Ca1A–O23–Ca1: 102.81(4)°. For clarity, hydrogen atoms bound to carbon are not shown.

To the best of our knowledge, such a structure has not been reported for a calcium pyrazolonato complex, but for the heavier analogs strontium(II) and barium(II) with four bridging ligands, instead of two, giving larger coordination numbers of 8 and higher.18 The structure contained a center of symmetry within the Ca2O2-ring. The bond lengths and angles are comparable to those found in literature for similar coordination environments.19

Transition Metal Complexes

Table 4 consists of a summary of selected bond lengths and angles of the crystal structures for the scandium-complexes 24 (Figure 5) and 25 (Figure 6) as well as the iron complex 46 (Figure 5). In the scandium(III)-complex 24, the scandium(III)-center is hepta-coordinated with three ligands and an additional water molecule, giving a distorted pentagonal bipyramidal structure. The values of selected bond lengths and angles for complex 24 correspond well to literature reported data on a structurally related fac-Sc-complex with an antipyrine-derived phenyl-protected pyrazolonato ligand (which bears a methyl group in the 3-position).20 The Sc—O(=C) bond lengths were determined to be 2.082(5), 2.106(5), and 2.104(4) Å, significantly shorter than the one in complex 24 (2.232(6), 2.249(5), and 2.239(5) Å). Presumably, the aquo ligand requires additional space, which shifts the pyrazolonato ligands, resulting in more distorted bond lengths and angles for the pseudo-octahedral complex. This also compares well to the structurally related archetype Sc(acac)3, where Sc–O bond lengths of 2.062(6)–2.082(6) Å have been determined, once more with the distinctive difference relative to the Sc—O(=C) bond lengths.21 On the other hand, scandium complex 25 formed a dimeric compound, consisting of two octahedral scandium centers bridged by two ethanoate ions to complete charge balance. The structure contains a core Sc2O2 ring with a C2 symmetry axis perpendicular to the ring, similar to the calcium complex 20. To the best of our knowledge, such a scandium(III)-pyrazolonato dimer has not been reported; however, related compounds include an aqua-malonato scandium(III) (Sc–O bond distances of 2.059(5) and 2.076(5) Å for the Sc2O2 ring; a Sc···Sc separation of 3.27 Å)22 as well as trinuclear scandium disiloxanediolate [{(Ph2Si–O)2O}2Sc3(acac)5], reported by Edelmann and co-workers (Sc–O(acac) bond distances of 2.095(2) Å, and Sc–O bond distances of 2.130(2) and 1.649(2) Å).23

Table 4. Selected Bond Lengths and Angles of the Scandium(II)-Pyrazolonato Complexes 24 and 25 as well as Iron(III)-Complex 46.
selected bond lengths [Å] and angles [°] 24 46 selected bond lengths [Å] and angles [°] 25
M–O3 2.062(5) 1.9761(15) M–O3/O3A 2.0885(15)
M–O12 2.232(6) 2.0815(18) M–O15/O15A 2.1297(14)
M–O23 2.096(5) 1.9484(18) M–O33/O33A 2.0960(15)
M–O32 2.249(5) 2.0679(17) M–O45/O45A 2.1106(14)
M–O43 2.144(4) 1.9548(16) M–O60 2.0705(14)
M–O52 2.239(5) 2.0500(15) M–O60A 2.0698(14)
M–O60 2.183(4)   Sc1–Sc1A 3.2508(7)
C–O3 1.307(9) 1.286(3) C–O3/O3A 1.281(3)
C–O12 1.274(9) 1.254(3) C–O15/O15A 1.265(3)
C–O23 1.295(8) 1.300(3) C–O33/O33A 1.282(2)
C–O32 1.265(8) 1.252(3) C–O45/O45A 1.268(2)
C–O43 1.275(8) 1.299(3) C–O60 1.422(4)
C–O52 1.259(8) 1.254(3) C–O60A 1.433(15)
O3–M–O12 83.0(2) 88.49(7) O3–M–O15 85.06(6)
O23–M–O32 78.6(2) 91.01(7) O33–M–O45 83.18(6)
O43–M–O52 81.39(18) 89.80(6) O60–M–O60A 76.53(6)
O3–M–O60 106.0(2)   Sc1–O60–Sc1A 103.47(6)
Figure 5.

Figure 5

Molecular structures of Sc-complex 24 and Fe-complex 46.

Figure 6.

Figure 6

Molecular structure of Sc-complex 25.

X-Ray structure determination of complex 46 revealed octahedral coordination of the central iron(III) atom, surrounded by three pyrazolonato ligands. In this arrangement, the complex is assigned as the Λ-isomer. The values reported in Table 4 agree well with those found in literature,24 particularly the slight difference in average bond length of the pyrazolonato–O–Fe bond of 1.9597(18) Å, being slightly shorter, than the average ester Fe—O(=C) bond of 2.066(18) Å. The Fe–O bonds are shorter than the corresponding Sc–O bonds found in complexes 24 and 25 as well as the literature-reported antipyrine-derived scandium(III)-complex.20a

The zirconium complexes 33 and 34 were found to be 8-coordinate, with four ligands around the zirconium(IV)-center, either in an antiprismatic or distorted antiprismatic arrangement (Figure 7). Consequently, those zirconium-pyrazolonato complexes of antiprismatic arrangement contained a center of symmetry. To the best of our knowledge, no structural data have been reported for a zirconium tetrapyrazolonato complex, with spectral data reported for a mixed zirconium pyrazolonato phthalocyaninate complex.25 Selected bond lengths and angles can be found in Table 5. The average Zr–O bond length is 2.207(2) Å, with a shortening noted along the Zr–O pyrazolonato bond (2.100(2) Å), and an elongation along the Zr—O(=C) bond (2.313(2) Å). The determined average Zr–O bond lengths of 2.207(2) Å compares well to the one found for Zr(acac)4 (2.188(1) Å).26

Figure 7.

Figure 7

General structure and atom labeling in Zr-complexes 33 and 34 listed in Table 5.

Table 5. Selected Bond Lengths and Angles of the Zirconium(IV)-Pyrazolonato Complexes.
selected bond lengths [Å] and angles [°] 33 34
M–O1 2.100(2) 2.0878(13)
M–O2 2.312(2) 2.2982(13)
M–O3 2.107(2) 2.0879(13)
M–O4 2.273(2) 2.2983(13)
M–O5 2.093(2) 2.0941(13)
M–O6 2.305(2) 2.3135(13)
M–O7 2.103(2) 2.0941(13)
M–O8 2.365(2) 2.3135(13)
C–O1 1.303(4) 1.289(2)
C–O2 1.243(4) 1.240(2)
C–O3 1.288(4) 1.289(2)
C–O4 1.239(4) 1.240(2)
C–O5 1.293(4) 1.282(2)
C–O6 1.239(5) 1.242(2)
C–O7 1.296(4) 1.282(2)
C–O8 1.235(5) 1.242(2)
O1–M–O2 79.59(9) 76.90(5)
O3–M–O4 79.54(9) 76.90(5)
O1–M–O3 148.78(9) 102.97(5)
O5–M–O6 77.95(9) 77.52(5)
O7–M–O8 78.03(9) 77.52(5)

The zirconium(IV) half-sandwich complexes 37 and 38, derived from zirconocene dichloride, showed similar Zr–O bond lengths as the complexes 33 and 34. Selected bond lengths and angles are summarized in Table 6. While complex 37 showed a monomeric structure with two pyrazolonato ligands, one cyclopentadienyl ligand and a chloride, complex 38 is dimeric having lost both chloride ions and being coordinated by four pyrazolonato ligands (Figure 8). The two units were found to be bridged by an oxide ligand, presumably from adventitious water in the reaction medium. The loss of a cyclopentadienyl ligand, presumably as cyclopentadiene, in these reactions most likely resulted by proton transfer from the ligand.27

Table 6. Selected Bond Lengths and Angles of the Zirconocene-Pyrazolonato Complexes.
selected bond lengths [Å] and angles [°] 37 selected bond lengths [Å] and angles [°] 38
M–O3/3A 2.1136(12) M–O3/3A 2.137(2)
M–O15/15A 2.2323(12) M–O12/12A 2.280(2)
M–CpC21 2.469(4) M–O23/23A 2.139(2)
M–CpC22 2.478(18) M–O32/32A 2.293(2)
M–CpC23 2.495(4) M–O50 1.9406(3)
M–CpC24 2.530(4) M–CpC41/41A 2.562(3)
M–CpC25 2.497(4) M–CpC42/42A 2.552(3)
M–Cl1 2.514(4) M–CpC43/43A 2.549(3)
C–O3 1.296(2) M–CpC44/44A 2.526(3)
C–O15 1.246(2) M–CpC45/45A 2.526(3)
CpC21–CpC22 1.42(2) C–O3/3A 1.277(3)
CpC22–CpC23 1.408(18) C–O12/12A 1.240(4)
CpC23–CpC24 1.387(6) C–O23/23A 1.291(3)
CpC24–CpC25 1.407(7) CpC41–CpC42 1.387(5)
CpC25–CpC21 1.401(8) CpC42–CpC43 1.415(5)
O3–M–O15 80.25(4) CpC43–CpC44 1.395(5)
O3–M–O15A 76.65(5) CpC44–CpC45 1.400(5)
O3–M–Cl1 94.53(14) CpC45–CpC41 1.399(5)
CpC21–CpC25–CpC24 107.9(4) O3–M–O12 79.72(7)
CpC23–CpC22–CpC21 105.6(14) O23–M–O32 79.31(8)
Figure 8.

Figure 8

Molecular structures of zirconium half sandwich complexes 37 and 38.

Moving along the 4d transition metal series, the rhodium centers were of slightly distorted octahedral coordination, with four acetate oxygens forming the square-planar arrangement, and the pyrazolone ligand in the axial position bonded via the nitrogen (Figure 9). The Rh–Rh-bond completed the octahedron. Selected bonds lengths and angles of complexes 39 and 40 are summarized in Table 6. The values compare well to those found in the rhodium acetate dimer bis-pyridine adduct (Rh–Rh bond length: 2.3963(2) Å, Rh–N bond length: 2.227(3) Å).28

Figure 9.

Figure 9

Molecular structures of rhodium-complexes 39 and 40.

The structure of the phenyl-protected pyrazolonato manganese(II) complex 42 was determined by X-ray crystallography, which showed a slightly distorted octahedral manganese(II)-bis(acylpyrazolonato) complex with a cis-configuration of the ligands (Figure 10), contrary to all previously reported pyrazolonato-Mn X-ray structures, with trans-stereochemistry of the ligands.12f,29 Selected bond lengths and angles are summarized in Table 8, which are in good agreement with those found in literature for the trans-configured manganese(II)-pyrazolonato complexes.12f,29

Figure 10.

Figure 10

Molecular structure of Mn-complex 42.

Table 8. Selected Bond Lengths and Angles of the cis-Configured Manganese(II)-, Zinc(II)- and Nickel(II)-Complexes.
selected bond lengths [Å] and angles [°] 42 50 Selected bond lengths [Å] and angles [°] 58
M–O3 2.1013(19) 1.992(3) M–O3 2.050(2)
M–O12 2.210(2) 2.146(3) M–O12 2.080(3)
M–O23 2.1029(18) 1.969(3) M–O23 2.025(3)
M–O32 2.220(2) 2.169(3) M–O32 2.107(3)
M–O40 2.196(2) 2.123(4) M–N40 2.084(3)
M–O50 2.201(2) 2.139(3) M–N50 2.071(3)
C–O3 1.268(3) 1.277(5) C–O3 1.272(4)
C–O12 1.238(3) 1.277(6) C–O12 1.250(5)
C–O23 1.265(3) 1.272(6) C–O23 1.285(5)
C–O32 1.244(3) 1.246(6) C–O32 1.236(5)
O3–M–O12 86.94(8) 90.46(14) O3–M–O12 92.35(1)
O23–M–O32 86.58(8) 90.18(13) O23–M–O32 90.33(11)
O40–M–O50 93.25(8) 92.31(13) N40–M–N50 91.41(12)
O3–M–O32 95.23(8) 92.52(13) O3–M–O32 87.88(10)

The structure of the phenyl-pyrazolone zinc(II) complex 50 showed the complex to have a cis-configuration (Figure 11), Table 7. Previous examples of zinc pyrazolonato complexes were reported to have trans relationship of the ligands and solvent molecules. In the cis-arrangement, the Zn–O bond lengths per pyrazolonato ligand were comparable to the ones of complex 49, see Table 8. Furthermore, the values compare well to those found in literature of related trans-arranged pyrazolonato-zinc complexes of octahedral coordination.8j,30 Particularly in a related cis-configured zinc(II)-pyrazolonato complex with a N,N,N′-trimethyl-1,2-ethylenediamine ligand (2.046(2) and 2.115(2) Å, and 2.054(2) and 2.115(2) Å, per pyrazolonato ligand; 2.150(2) and 2.192(2) Å for the diamine ligand)31 as well as a cis-configured zinc(II)-pyrazolonato complex with 1,10-phenanthroline (2.0347(13) and 2.1123(14) Å per pyrazolonato ligand; 2.1908(16) Å for Zn–N(phen)),32 the bond lengths were found to be in agreement with those of complex 50. Zinc(II)-pyrazolonato complexes of square-planar and square-pyramidal geometry were found to exhibit similar Zn–O bond lengths.8j,32,33

Figure 11.

Figure 11

Molecular structures of Zn- and Ni-complexes 50 and 58, respectively.

Table 7. Selected Bond Lengths and Angles of the Dirhodium(II)-Pyrazolonato Complexes.
selected bond lengths [Å] and angles [°] 39 selected bond lengths [Å] and angles [°] 40
M–O20/20A 2.031(2) M–O20/20A 2.037(2)
M–O22/22A 2.039(2) M–O22/22A 2.039(2)
M–O30/30A 2.049(2) M–O30/30A 2.045(2)
M–O32/32A 2.034(2) M–O32/32A 2.033(2)
M–N1 2.278(3) M–N1 2.276(2)
Rh1–Rh1A 2.3956(5) Rh1–Rh1A 2.4015(5)
C–O3 1.349(4) C–O3 1.328(4)
C–O15 1.224(5) C–O12 1.220(4)
C21–O20/22 1.259(4) C21–O20/22 1.266(4)/1.262(4)
C31–O30/32 1.263(4)/12.64(4) C31–O30/32 1.266(4)/1.263(4)
O20–M–O30 88.69(10) O20–M–O30 87.09(9)
O20–M–N1 93.53(10) O20–M–N1 90.90(9)
O20–M–O22A 176.09(10) O20–M–O22A 175.70(9)
O20–M–O32A 91.35(1) O20–M–O32A 91.21(10)
O22–M–O30 90.61(10) O22–M–O30 91.61(10)
O22–M–O32 89.07(10) O22–M–O32 89.79(10)

Crystals of the nickel(II) complex 58 suitable for X-ray crystallographic analysis were obtained after recrystallization from pyridine. Complex 58 is octahedral with the two pyrazolonato ligands in a cis-arrangement, Figure 11. The structures of related compounds display Ni–O bond lengths of 1.9966(12)–2.0442(17) Å, while the solvent ligands were reported with bond distances to the Ni(II) center of 2.0651(16)–2.042(12) Å.12

The zinc complex 49 with the PMB-protected ligand has a polymeric structure in the solid state (Figure 12). The zinc(II) center is octahedrally coordinated, with the pyrazolonato ligand in the plane, while the axial positions were occupied by nitrogen atoms of ligands on neighboring molecules. This polymeric structure was evident in the solution as well in that attempts to record 1H-NMR spectra on complex 59 in solvents such as CDCl3, CD3OD, CD3CN, and (CD3)2CO all failed, and only when dissolved in (CD3)2SO was it possible to obtain NMR data. This only showed the heterocyclic bidentate ligand, with no other ligands present, further confirming the absence of any solvent molecules in the complex. Similar NMR behavior was observed for the benzoyl-substituted pyrazolonato complex 51, although no X-ray crystal structure determination could be obtained for this complex.

Figure 12.

Figure 12

Crystal structure of Zn-complex 49.

The zinc(II) ion in complex 49 possessed a slightly distorted octahedral coordination, with two sites occupied by the pyrazolonato ligands, and the other two coordination sites were occupied by the pyrazolonato-nitrogen atoms (cf. the sodium complex 14). Selected bond lengths and angles are summarized in Table 9. The values are in good agreement with the zinc(II) complex 50 and compare well to those reported in the literature.8j,3033

Table 9. Selected Bond Lengths and Angles of the Polymeric Zinc(II) and Copper(II)-Complexes.
selected bond lengths [Å] and angles [°] 49 selected bond lengths [Å] and angles [°] 54 selected bond lengths [Å] and angles [°] 55
M–O3/3A 2.0737(13) M1A–O3/3A 1.927(2) M1A–O3/33 1.948(4)
M–O15/15A 2.1381(13)/2.1382(13) M1A–O12/12A 2.001(2) M1A–O15/45 1.985(4)/2.000(4)
M–N1B/1D 2.1719(16) M1A–N1B/1D 2.519(3) M1A–N1A/31C 2.458(4)/2.626(4)
C–O3 1.272(2) C–O3 1.275(4) C–O3 1.271(6)
C–O15 1.235(2) C–O12 1.249(4) C–O12 1.268(6)
O3–M–O15 92.40(5) O3–M–O12 94.59(9) O3–M–O15 94.83(15)
O3A–M–O15 87.60(5) O3A–M–O12 85.41(9) O33–M–O15 85.58(16)
O3–M–N1B 91.19(6) O3–M–N1B 94.44(10) O3–M–N1A 84.12(15)
O3A–M–N1B 88.81(6) O3A–M–N1B 85.56(10) O33–M–N31C 89.02(16)
O15–M–N1B 86.78(6) O12–M–N1B 92.37(9) O15–M–N1A 82.05(14)
O15A–M–O1B 93.22(6) O15A–M–N1B 87.63(9) O15–M–N31C 97.17(15)

The structurally related phenyl-protected pyrazolone copper(II)-complex 54 and benzoyl substituted pyrazolone copper(II) complex 55 were characterized after recrystallization from ethanol and ethanol/pyridine/DMF, respectively, and were found to have a polymeric structure in the solid state (Figure 13). The central copper(II) ions are octahedrally coordinated, with the pyrazolonato ligand in the horizontal plane, while the axial positions are occupied by nitrogen atoms of neighboring ligand molecules, Table 9. Cu–N bond lengths of 2.519(3) Å in complex 54 and 2.458(4)/2.626(4) Å for complex 55 are consistent with Jahn-Teller distortion, with shortened Cu–O bond lengths of 1.927(2) and 2.001(2) Å (complex 54) and 1.948(4)/1.985(4)/2.000(4) Å (complex 55). Goetz-Grandmont et al. reported the crystal structure of a square-planar copper-pyrazolonato complex with two centrosymmetric molecules in the asymmetric unit.34a These showed comparable Cu–O bond lengths of 1.912(8)–1.939(8) Å. Furthermore, Jahn-Teller distorted bonding between the two units was observed, displaying a Cu–O bond length of 2.546(8) Å.

Figure 13.

Figure 13

Crystal structures of Cu-complexes 54 and 55.

Crystals of the nickel(II) complexes 60 (Figure 14) and 57 (Figure 15) were obtained after recrystallization from pyridine. Crystallographic analysis revealed that the nickel(II) complexes were of slightly distorted octahedral coordination with the two pyrazolonato ligands in a trans-arrangement, Table 10. The reported values correspond reasonably well to reported data with structures, which have been shown to display Ni–O bond lengths of 1.9966(12)–2.0442(17) Å, while the solvent ligands were reported with bond distances to the Ni(II) center of 2.0651(16)–2.042(12) Å.12

Figure 14.

Figure 14

Molecular structures of Cu- and Ni-complexes 56 and 60.

Figure 15.

Figure 15

Molecular structure of Ni-complex 57.

Table 10. Selected Bond Lengths and Angles of Pyridine-Coordinated Nickel(II) and Copper(II) Complexes.
selected bond lengths [Å] and angles [°] 56 selected bond lengths [Å] and angles [°] 60 selected bond lengths [Å] and angles [°] 57
M–O3/3A 1.9790(13) M–O3/13 2.0613(16)/2.0686(15) M–O3/23 1.96(3)/2.11(3)
M–O6/6A 2.3783(13) M–O6/16 2.0726(15)/2.0785(15) M–O15/35 1.090(16)/2.052(16)
M–N11/11A 2.0310(16)/2.0311(16) M–N21/31 2.088(2)/2.0895(19) M–N41/51 2.16(2)/2.01(2)
C–O3 1.295(2) C–O3/13 1.284(3)/1.282(3) C–O3/23 1.267(11)/1.266(10)
C–O6 1.225(2) C–O6/16 1.243(3)/1.246(3) C–O15/35 1.238(12)/1.249(10)
O3–M–O6 88.99(5) O3–M–O6 93.65(6) O3–M–O15 96.0(9)
O3A–M–O6 91.01(5) O13–M–O16 94.11(6) O23–M–O35 92.2(9)
O3–M–N11 92.14(6) O13–M–O6 86.97(6) O23–M–O15 86.0(8)
O3A–M–N11 87.86(6) O3–M–N21 89.94(7) O3–M–N41 90.2(13)
    O13–M–N21 89.05(7) O23–M–N21 92.4(12)

Recrystallization of the copper(II) complexes 53 and 56 (Figure 14) from pyridine gave crystals suitable for X-ray crystal structure determinations. These were consistent with slightly distorted octahedral coordination with two pyridine solvent molecules trans to each other in the remaining coordination sites. Complex 56 was found to be Jahn-Teller distorted along the equatorial pyrazolone Cu–O bond with a Cu–O bond length of 2.3783(13) Å. The values reported in Table 10 compare well to a known antipyrine-derived copper complex with two phenyl-protected pyrazolonato ligands and two pyridine ligands.34c Related copper(II) pyrazolonato complexes have been reported with distorted octahedral coordination33b,34b,34c as well as square-planar,8f,12f,33b,34a or square-pyramidal cores.16b In these octahedral complexes, Cu–O bonds lengths were reported to be 1.953(3)–1.979(3) and 2.259(4)–2.383(4) Å with distortion along the C—O(=C) bond and with additional donor ligands (pyridine, bipy, phen) at a distance of 2.004(2)–2.0308(16) Å. Dey et al. reported a copper(II)-pyrazolonato complex with two methanol ligands that showed two shorter bonds of 2.033(3) Å, and four longer Cu–O bonds at 2.108(4)–2.152(4) Å. The square-planar complexes were reported with Cu–O bond lengths of 1.907(8)–1.939(8) Å, while the square-pyramidal complex showed Cu–O bond lengths of 1.911(3)–1.956(3) Å with Jahn-Teller distorted Cu–O bonds of 2.248(4) Å.

This unusual discrepancy of Cu–O bond elongation over Cu–N bond elongation in complex 56, while the polymeric complexes 54 and 55 showed Cu–N elongation over Cu–O elongation are noteworthy. In addition, literature reports Jahn-Teller distorted Cu–O intermolecular binding instead of C–N binding. Before undertaking DFT calculations (at the M062X/Def2-SVP level)50 to provide an insight into the Jahn-Teller distorted bond lengths of the synthesized complexes, we examined their statistical distribution in octahedral copper(II) complexes exhibiting four di-axial Cu–O bonds and two di-axial Cu–N bonds, as found in the Cambridge Structural Database (CSD). This revealed (April, 2023) 1412 error-free structures where the shortened Jahn-Teller distorted O–Cu(II) bonds are ∼1.9 Å and the elongated Jahn-Teller distorted O–Cu bonds are 2.4–2.6 Å, as in complex 56. The most probable distribution (SI Figure S107) is where one pair of O–Cu bonds is ∼0.6 Å longer than the other, but with a significant, albeit smaller and compact, distribution where both pairs of O–Cu bonds have the same length of ∼1.9 Å and yet another where they both have a length of ∼2.2 Å. A plot of the mean of N–Cu–N lengths vs the mean of O–Cu–O lengths (SI Figure S108) shows a hot spot of both short N–Cu lengths and O–Cu lengths, (that observed for, e.g., complex 56) with an accompanying but more diffuse distribution with short N–Cu lengths and long O–Cu lengths. Significantly fewer structures were found that displayed shortened O–Cu bonds with longer Cu–N bonds, as in complex 54, suggesting that the O–Cu elongated case in complex 56 is statistically more commonly found in the literature. Correlations of these bond lengths to bond strengths have been reported.35

The DFT calculations matched this observation, with a starting geometry containing a N–Cu bond-elongated isomer of complex 56 collapsing without activation on optimization to the N–Cu shortened isomer. The form of the SOMO is dominated by the antibonding interaction of the dx2y2 orbital on copper(II) with the lone pair of the pyridine nitrogen and a lone pair on the pyrazolonato oxygen in a square-planar arrangement (Figure 16).

Figure 16.

Figure 16

SOMO (iso-surface value is 0.02au) of copper(II) complex (56).

Lanthanide/Rare Earth and Actinide Complexes of Acylpyrazolone Ligands

The yttrium(III)-ion in complex 28 was found to be eight-coordinate, with three acylpyrazolonato ligands and two aquo ligands (Figure 17). Selected bond lengths and angles of complex 28 are reported in Table 11. These values compare well to literature reported data,36 with a slight elongation noted along the Y—O(=C) bond relative to literature data (2.4099(18) Å vs 2.282(2) Å/36a 2.321(6) Å36b).

Figure 17.

Figure 17

Molecular structures of Y- and La-complexes 28 and 62, respectively.

Table 11. Selected Bond Lengths and Angles of the Isomorphous Ytterbium(III)-, Lanthanum(III)-, Dysprosium(III)-, and Yttrium(III)-Phenyl-Pyrazolone Complexes.
selected bond lengths [Å] and angles [°] 28 62 66 70
M–O3 2.2899(18) 2.4734(14) 2.253(3) 2.258(2)
M–O12 2.4026(18) 2.5234(15) 2.433(3) 2.370(2)
M–O23 2.29(52(18) 2.4117(14) 2.374(3) 2.258(2)
M–O32 2.3933(18) 2.5803(14) 2.392(3) 2.3601(19)
M–O43 2.2992(19) 2.4673(14) 2.331(3) 2.267(2)
M–O52 2.4340(19) 2.5437(15) 2.436(3) 2.418(2)
M–O60 2.3656(18) 2.5609(16) 2.390(3) 2.345(2)
M–O70 2.3761(19) 2.4910(14) 2.358(3) 2.356(2)
C–O3 1.268(3) 1.264(3) 1.288(5) 1.274(4)
C–O12 1.239(3) 1.238(2) 1.233(5) 1.246(4)
C–O23 1.273(3) 1.273(2) 1.259(5) 1.271(4)
C–O32 1.237(3) 1.234(2) 1.242(5) 1.242(4)
C–O43 1.277(3) 1.266(2) 1.257(5) 1.274(4)
C–O52 1.236(3) 1.229(3) 1.238(5) 1.238(4)
O3–M–O12 74.12(6) 71.73(5) 76.47(10) 74.97(7)
O23–M–O32 74.62(6) 73.23(5) 75.17(10) 75.56(7)
O43–M–O52 77.48(7) 69.94(5) 73.23(11) 78.37(8)
O60–M–O70 137.84(6) 89.93(5) 9785(10) 137.46(7)

The lanthanum(III)-phenyl pyrazolonato complex 62 is isostructural to yttrium complex 28, while the solvent molecules L were H2O and EtOH (Figure 17). Selected bond lengths and angles are reported in Table 11. These values compare reasonably well to known lanthanum(III)-pyrazolonato complex with two additional water molecules of the general formula [La(QPh)3(H2O)2], in which La–O bond distances for the ligand were reported to be between 2.431 and 2.518 Å, while the aquo ligands were determined to be 2.530 and 2.617 Å.13 The values in complex 62 are, furthermore, in good agreement with a known formyl-substituted dimeric lanthanum(III)-pyrazolonato complex, which displayed similar La–O bond lengths between 2.436(3) and 2.650(4) Å and aquo ligands at a La–O distance of 2.560(3)–2.676(3) Å.37

X-Ray crystal structure determination of the dysprosium(III)-complex 66 revealed it to be isomorphous to the lanthanum(III) complex 62 (Figure 18). Consistent with lanthanide contraction, the Dy–O bond lengths were found to be generally shorter than the La–O bond lengths. In the solid state, the phenyl protected pyrazolonato dysprosium complex 66 gave an octacoordinated dysprosium(III) center, with two positions occupied by aquo ligands. Selected bond lengths and angles are reported in Table 11. Comparable Dy–O bond lengths (2.306(2)–2.380(2) Å) and angles (74.86(8)°–75.49(8)°) have been reported by Jia et al.,38 Pettinari et al. (Dy–O bond length between 2.271(3) and 2.410 Å),39 Li et al. (2.335(3)–2.422(3) Å),40 and Yu et al. (Dy–O bond length between 2.261(3)–2.492(3) Å).41

Figure 18.

Figure 18

Molecular structures of Dy- and Yb-complexes 66 (L = H2O) and 70 (L = H2O), respectively.

In the solid state, the phenyl-protected pyrazolonato ytterbium(III) complex 70 was isomorphous with complexes 28, 62, and 66 (Figure 18). The Yb–O bond lengths were found to be shorter than the La–O and Dy–O bond lengths, consistent with the lanthanide contraction. The ytterbium(III) center is octacoordinated with two additional aquo ligands completing the coordination sphere. Selected bond lengths and angles are reported in Table 11. Comparable Yb–O bond lengths (2.255(4)–2.360(5) Å) and angles (74.6(2)°–75.2(2)°) in ytterbium pyrazolonato complexes with solvent molecules (H2O or EtOH) have been reported by Bombierie et al.,42 and Yang and co-workers (2.257–2.370 Å),29 while Lü and co-workers reported structural data on a mixed ytterbium(III)-pyrazolonato complex with 1,10-phenanthroline (Yb–O bond lengths of: 2.252(7)–2.369(8) Å;O–Yb–O angles per ligand of: 71.9(3)°–73.6(3)°).43 Lü et al. further reported Yb–O bond lengths of similar values (2.261(5)–2.399(5) Å) in mixed Zn-salen Yb-tris pyrazolonato complexes with slightly smaller dihedral angles per pyrazolonato ligand (72.26(15)°–73.03(18)°).43

The reactions with the PMB-protected pyrazolone 8 and lanthanum(III) or dysprosium(III) triflate gave crystals of the coordination polymers 61 and 65, which were isomorphous. The lanthanide centers in both compounds are octa-coordinated, however with four acylpyrazolonato ligands, giving a formal negative charge, balanced with a sodium cation (Figure 19). The pyrazolonato ligands were determined to be symmetry equivalent, and the sodium cation is coordinated to the pyrazolone nitrogens. Selected bond lengths and angles are depicted in Table 12. To the best of our knowledge, a MOF-like complex of lanthanum or dysprosium with four acylpyrazolone ligands and a sodium cation has not been previously reported in literature. In the mixed lanthanum/sodium-polymer, per unit cell, two cavities were found in the MOF, with volumes of 133 Å3, which were found to partially host recrystallization solvent molecules. Similarly in the mixed dysprosium(III)/sodium(I)-polymer, per unit cell, two cavities were found in the MOF, with volumes of 184 Å3, which were found to partially host recrystallization solvent molecules, Figure 20.

Figure 19.

Figure 19

Crystal structures of the heterobimetallic sodium/lanthanide coordination polymers units 61 and 65.

Table 12. Selected Bond Lengths and Angles of Heterobimetallic Sodium/Lanthanide Coordination Polymers.
selected bond lengths [Å] and angles [°] 61 65
M–O3/3A/3B/3C 2.401(12)/2.402(12) 2.289(7)
M–O15/15A/15B/15C 2.546(17) 2.455(6)
Na–N1/1D/1E/1F 2.357(19) 2.425(9)
C–O3 1.319(18) 1.275(11)
C–O15 1.28(3) 1.233(11)
O3–M–O15 74.4(6) 76.1(2)
N1–M–N1D 90.9(8) 89.1(5)
N1–M–N1E 119.5(5) 120.5(2)
Figure 20.

Figure 20

Illustration of the network of MOF-polymer 65 (color scheme: C, gray; O, red; N, dark purple; Dy, turquoise; Na, light purple).

Crystal structures of polymeric lanthanide acylpyrazolone complexes have been reported for samarium(III), europium(III), terbium(III), and dysprosium(III) with hydroxonium counter-cations.44 In addition, Shul’gin et al. reported the X-ray structure determination of a MOF-like acylpyrazolonato complex with europium(III), and sodium counter-cations.45 A related complex with a significantly larger I-N-dodecyl-N′,N′-dimethylamino-stilbazolium counterion has been reported, but this one has been determined to display discrete molecules in the solid state, which are not polymeric.46 In the octacoordinated lanthanum(III)-complex anion, La–O bond lengths of 2.454(4)–2.530(3) Å were determined, with O–La–O bond angles per pyrazolonato ligand of 69.16(9)°–70.52(9)°, values which are in good agreement with the ones determined for polymeric complex (61). X-Ray structures of related nonpolymeric lanthanide acylpyrazolone complexes with four ligand molecules and ammonium-derived counter-cations have been reported.39,45,47 Related ammonium complexes have been reported in studies of dysprosium(III) pyrazolonato complexes as constituents of Langmuir–Blodgett films and for other material applications.47a,48

The uranium(VI) ion in complexes 73 (Figure 21), 74 and 76 (Figure 22), and 75 (Figure 23) were found to contain a distorted octahedral core structure with two oxido ligands making up the uranyl(VI) ion, coordinated by two acylpyrazolonato ligands and an additional solvent ligand (methanol, ethanol or pyridine) resulting in the overall pentagonal-bipyramidal structure of the hepta-coordinated complexes. Complexes 73, 75, and 76 were found to contain a mirror symmetry element along the UO2–solvent ligand plane, while complex 74 was found to lack this symmetry element with the two acylpyrazolonato ligands being in trans-relationship to each other. Selected bond lengths and angles for complexes 7376 are summarized in Tables 13 and 14. These values compare well to literature reported data for related uranyl acylpyrazolonato complexes.49

Figure 21.

Figure 21

Molecular structure of UO2-complex 73.

Figure 22.

Figure 22

Molecular structures of UO2-complexes 74 and 76.

Figure 23.

Figure 23

Molecular structures of UO2 complex 75 with pyridine (A) and methanol (B) as ligands.

Table 13. Selected Bond Lengths and Angles of Uranyl Acylpyrazolonato Complexes 73, 74, and 76.
selected bond lengths [Å] and angles [°] 73 selected bond lengths [Å] and angles [°] 74 selected bond lengths [Å] and angles [°] 76
M–O50/60 1.765(16)/1.785(17) M–O50/60 1.731(10)/1.736(11) M–O1/2 1.773(3)/1.763(4)
M–O3/23 2.321(13)/2.361(13) M–O3/23 2.335(7)/2.297(9) M–O3/5 2.297(3)/2.359(3)
M–O15/35 2.454(13)/2.458(13) M–O12/32 2.430(10)/2.452(8) M–O4/6 2.411(3)/2.424(3)
M–O40 2.368(15) M-N41 2.551(11) M–N5 2.557(4)
C–O3/23 1.27(2)/1.29(3) C–O3/23 1.279(13)/1.268(15) C–O3/5 1.290(6)/1.297(6)
C–O15/35 1.26(2)/1.27(2) C–O12/32 [1.19(3) + 1.36(3)]/1.244(15) C–O4/6 1.239(6)/1.238(6)
O3–M–O15/O23–M–O35 72.9(5)/73.8(5) O3–M–O12/O23–M–O32 72.5(3)/73.2(3) O3–M–O4/O5–M–O6 73.62(12)/74.03(11)
O50–M–O60 179.2(9) O50–M–O60 178.0(4) O1–M–O2 176.14(13)
O50–M–O40/O60–M–O40 89.5(6)/90.6(7) O50–M–N41/O60–M–N41 89.4(4)/88.6(4) O1–M–N5/O2–M–N5 88.81(14)/87.36(15)
Table 14. Selected Bond Lengths and Angles of Uranyl Acylpyrazolonato Complex 75.
selected bond lengths [Å] and angles [°] 75-A selected bond lengths [Å] and angles [°] 75-B
M–O70A/80A 1.750(8)/1.763(8) M–O70B/80B 1.758(7)/1.768(7)
M–O3A/33A 2.388(7)/2.358(6) M–O3B/33B 2.358(8)/2.347(7)
M–O15A/45A 2.367(6)/2.369(7) M–O15B/45B 2.393(7)/2.416(9)
M–N60A 2.568(8) M-O60B 2.391(5)
C–O3A/33A 1.283(11)/1.289(12) C–O3B/33B 1.266(12)/1.289(13)
C–O15A/45A 1.269(11)/1.271(11) C–O15B/45B 1.269(12)/1.284(12)
O3A–M–O15A/O33A–M–O45A 73.3(2)/73.1(2) O3B–M–O15B/O33B–M–O45B 72.2(3)/72.6(3)
O70A–M–O80A 179.1(3) O70B–M–O80B 179.2(4)
O70A–M–N60A/O80A–M–N60A 88.0(3)/91.3(3) O70B–M–O60B/O80B–M–O60B 95.8(3)/84.7(3)

Conclusions

In conclusion, structurally diverse pyrazolone-metal complexes were synthesized from readily accessible pyrazolone precursors. The complexes contained the main group and transition metals, lanthanides, and actinides. Structural characterization was performed by NMR and IR spectroscopy, X-ray crystal structure determinations, mass-spectrometry, and DSC measurements. Multiple examples were found to be of a polymeric structure in the solid state—a feature which was observed on complexes of all four ligands. Future investigations are directed toward the material properties of the MOF-like polymers and the extension toward other metal ions of interest.

Acknowledgments

Peter Haycock, Dr. Stuart Elliot and Dr. Lisa Haigh (Imperial College London) are gratefully acknowledged for their help with NMR and mass spectrometry, respectively. We also extend our thanks to Nigel Howard (University of Cambridge) and Orla McCullough (London Metropolitan University) for elemental microanalyses, as well as Prof. Stephen Skinner and Jiao Gao (Materials Department, Imperial College London) for support with TG/DSC and TG/DTA measurements. We thank GlaxoSmithKline for the generous endowment (to A.G.M.B.).

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.inorgchem.3c01506.

  • Spectral data, DSC and TG/DTA measurements, computational calculations, and X-ray crystallography data (PDF)

  • Detailed experimental procedure, compound characterization, and NMR-spectra of all synthesized compounds as well additional X-Ray crystallography data Raw computational and crystallographic FAIR data are available from the Spiral data repository via ref (50). (PDF)

The authors thank the Leverhulme Trust for a Research Project Grant (Ref No.: RPG-2020-273 awarded to Anthony G. M. Barrett) and the Engineering and Physical Sciences Research Council (EPSRC) for grants EP/V003089/1 and EP/V046659/1 (both awarded to Richard A. Layfield).

The authors declare no competing financial interest.

Supplementary Material

ic3c01506_si_001.pdf (44MB, pdf)
ic3c01506_si_003.pdf (9.5MB, pdf)

References

  1. a Lukehart C. M. Metalla-β-diketones and Their Derivatives. Acc. Chem. Res. 1981, 14, 109–116. 10.1021/ar00064a003. [DOI] [Google Scholar]; b Pettinari C.; Marchetti F.; Drozdov A.. 1.6 β-Diketones and Related Ligands. In Comprehensive Coordination Chemistry II; Editor(s): McCleverty J. A.; Meyer T. J.; Elsevier, Amsterdam, 2003; pp. 97–115, 10.1016/B0-08-043748-6/01181-6. [DOI] [Google Scholar]; c Skopenko V. V.; Amirkhanov V. M.; Yu Sliva T.; Vasilchenko I. S.; Anpilova E. L.; Garnovskii A. D. Various Types of Metal Complexes Based on Chelating β-diketones and Their Structural Analogues. Russ. Chem. Rev. 2004, 73, 737–752. 10.1070/RC2004v073n08ABEH000909. [DOI] [Google Scholar]; d Garnovskii A. D.; Vasil’chenko I. S. Tautomerism and various coordination modes of typical chelating ligands with metals. Russ. Chem. Rev. 2005, 74, 193–215. 10.1070/RC2005v074n03ABEH001164. [DOI] [Google Scholar]; e Mehrotra R. C. Chemistry of metal β-diketonates. Pure Appl. Chem. 1988, 60, 1349–1356. 10.1351/pac198860081349. [DOI] [Google Scholar]; f Vigato P. A.; Peruzzo V.; Tamburini S. The evolution of β-diketone or β-diketophenol ligands and related complexes. Coord. Chem. Rev. 2009, 253, 1099–1201. 10.1016/j.ccr.2008.07.013. [DOI] [Google Scholar]; g Kawaguchi S. Variety in the coordination modes β-dicarbonyl compounds in metal complexes. Coord. Chem. Rev. 1986, 70, 51–84. 10.1016/0010-8545(86)80035-2. [DOI] [Google Scholar]
  2. a Hagen H.; Boersma J.; van Koten G. Homogeneous vanadium-based catalysts for the Ziegler-Natta polymerization of α-olefins. Chem. Soc. Rev. 2002, 31, 357–364. 10.1039/B205238E. [DOI] [PubMed] [Google Scholar]; b Trzeciak A.; Borak B.; Ciunik Z.; Ziółkowski J.; Guedes da Silva M.; Pombeiro A. L. Structure, Electrochemistry and Hydroformylation Catalytic Activity of the Bis(pyrazolylborato)rhodium(I) Complexes [RhBp(CO)P] [P = P(NC4H4)3, PPh3, PCy3, P(C6H4OMe-4)3]. Eur. J. Inorg. Chem. 2004, 2004, 1411–1419. 10.1002/ejic.200300517. [DOI] [Google Scholar]; c Harris T.M. (2001). 2,4-Pentanedione. In Encyclopedia of Reagents for Organic Synthesis, (Ed.). 10.1002/047084289X.rp030. [DOI] [Google Scholar]
  3. a Nagai A.; Kokado K.; Nagata Y.; Chujo Y. 1,3-Diketone-Based Organoboron Polymers: Emission by Extending π-Conjugation along a Polymeric Ligand. Macromolecules 2008, 41, 8295–8298. 10.1021/ma801690d. [DOI] [Google Scholar]; b Otsuka T.; Chujo Y. Blue Emission from Polymer Nanocomposites: Preparation and Application of Multicolored Luminescent Materials. Polym. J. 2011, 43, 352–357. 10.1038/pj.2010.144. [DOI] [Google Scholar]; c Aromí G.; Gamez P.; Reedijk J. Poly β-diketones: Prime Ligands to Generate Supramolecular Metalloclusters. Coord. Chem. Rev. 2008, 252, 964–989. 10.1016/j.ccr.2007.07.008. [DOI] [Google Scholar]; d Brock A. J.; McMurtrie J. C.; Clegg J. K. Self-Assembly of bis-β-diketone-based [M2L2] Dinuclear Platforms into 2-Dimensional Coordination Polymers. CrystEngComm 2019, 21, 4786–4791. 10.1039/C9CE00896A. [DOI] [Google Scholar]; e Carlucci L.; Ciani G.; Proserpio D. M.; Visconti M. The novel metalloligand [Fe(bppd)3] (bppd = 1,3-bis(4-pyridyl)-1,3-propanedionate) for the crystal engineering of heterometallic coordination networks with different silver salts. Anionic control of the structures. CrystEngComm 2011, 13, 5891–5902. 10.1039/C1CE05520H. [DOI] [Google Scholar]; f Burrows A. D.; Mahon M. F.; Renouf C. L.; Richardson C.; Warren A. J.; Warren J. E. Dipyridyl β-diketonate complexes and their use as metalloligands in the formation of mixed-metal coordination networks. Dalton Trans. 2012, 41, 4153–4163. 10.1039/C2DT12115H. [DOI] [PubMed] [Google Scholar]; g Brock A. J.; Clegg J. K.; Li F.; Lindoy L. F. Recent developments in the metallo-supramolecular chemistry of oligo-β-diketonato ligands. Coord. Chem. Rev. 2018, 375, 106–133. 10.1016/j.ccr.2017.11.007. [DOI] [Google Scholar]; h Clegg J. K.; Li F.; Lindoy L. F. Oligo-b-diketones as versatile ligands for use in metallo-supramolecular chemistry: Recent progress and perspectives. Coord. Chem. Rev. 2022, 455, 2144355. 10.1016/j.ccr.2021.214355. [DOI] [Google Scholar]
  4. a Jiang S.-D.; Wang B.-W.; Su G.; Wang Z.-M.; Gao S. A Mononuclear Dysprosium Complex Featuring Single-Molecule-Magnet Behavior. Angew. Chem., Int. Ed. 2010, 49, 7448–7451. 10.1002/anie.201004027. [DOI] [PubMed] [Google Scholar]; b Williamson K. S.; Yoon T. P. Iron-Catalyzed Aminohydroxylation of Olefins. J. Am. Chem. Soc. 2010, 132, 4570–4571. 10.1021/ja1013536. [DOI] [PMC free article] [PubMed] [Google Scholar]; c Sudo A.; Hirayama S.; Endo T. Highly efficient catalysts-acetylacetonato complexes of transition metals in the 4th period for ring-opening polymerization of 1,3-benzoxazine. J. Polym. Sci., Part A: Polym. Chem. 2010, 48, 479–484. 10.1002/pola.23810. [DOI] [Google Scholar]; d Bray D. J.; Clegg J. K.; Lindoy L. F.; Schilter D. Self-assembled Metallo-supramolecular Systems Incorporating b-Diketone Motifs as Structural Elements. Adv. Inorg. Chem. 2006, 59, 1–37. 10.1016/S0898-8838(06)59001-4. [DOI] [Google Scholar]; e Crossman A. S.; Marshak M. P.. 1.11 β-Diketones: Coordination and Application. In Comprehensive Coordination Chemistry III; Editor(s): Constable E. C.; Parkin G.; Que L. Jr.; Elsevier, Amsterdam, 2021; pp. 331–365, 10.1016/B978-0-08-102688-5.00069-6. [DOI] [Google Scholar]; f Zhao Z.; Baba Y.; Kubota F.; Kamiya N.; Goto M. Synergistic Extraction of Rare-Earth Metals and Separation of Scandium Using 2-Thenoyltriuoroacetone and Tri-n-octylphosphine Oxide in an Ionic Liquid System. J. Chem. Eng. Jpn. 2014, 47, 656–662. 10.1252/jcej.14we360. [DOI] [Google Scholar]; g Zhang J.; Tanjedrew N.; Wenzel M.; Royla P.; Du H.; Kiatisevi S.; Lindoy L. F.; Weigand J. J. Selective Separation of Lithium, Magnesium and Calcium using 4-Phosphoryl Pyrazolones as pH-Regulated Receptors. Angew. Chem., Int. Ed. 2023, 62, e202216011 10.1002/anie.202216011. [DOI] [PubMed] [Google Scholar]; h Zhang J.; Wenzel M.; Schnaars K.; Hennesdorf F.; Lindoy L. F.; Weigand J. J. Highly Tunable 4-Phosphoryl Pyrazolone Receptors for Selective Rare-Earth Separation. Inorg. Chem. 2023, 62, 3212–3228. 10.1021/acs.inorgchem.2c04221. [DOI] [PubMed] [Google Scholar]
  5. a Marchetti F.; Pettinari C.; Pettinari R. Acylpyrazolone ligands: Synthesis, structures, metal coordination chemistry and application. Coord. Chem. Rev. 2005, 249, 2909–2945. 10.1016/j.ccr.2005.03.013. [DOI] [Google Scholar]; b Marchetti F.; Pettinari C.; Pettinari R. Recent advances in acylprazolone metal complexes and their potential applications. Coord. Chem. Rev. 2015, 303, 1–31. 10.1016/j.ccr.2015.05.003. [DOI] [Google Scholar]; c Marchetti F.; Pettinari C.; Di Nicola C.; Tombesi A.; Pettinari R. Coordination chemistry of pyrazolone-based ligands and application of their metal complexes. Coord. Chem. Rev. 2019, 401, 213069 10.1016/j.ccr.2019.213069. [DOI] [Google Scholar]
  6. a Jensen B. S.; Sörensen N. A.; Jensen C. E. Solvent Extraction of Metal Chelates. II. An Investigation of some 1-Phenyl-3-methyl-4-acyl-pyrazolones-5. Acta Chem. Scand. 1959, 13, 1890–1896. 10.3891/acta.chem.scand.13-1890. [DOI] [Google Scholar]; b Zolotov Y. A.; Kuzmin N. M.. Extraction of Metals by Acylpyrazolones; Mauka: Moscow, Russia, 1977. [Google Scholar]; c Rao G. N.; Arora H. C. Solvent extraction of uranium(VI) with 4-acyl-2,4-dihydro 5-methyl 2-phenyl 3H-pyrazol 3-ones. J. Inorg. Nucl. Chem. 1977, 39, 2057–2060. 10.1016/0022-1902(77)80546-0. [DOI] [Google Scholar]; d Sasaki Y.; Freiser H. Mixed-ligand chelate extraction of lanthanides with 1-phenyl-3-methyl-4-acyl-5-pyrazolones. Inorg. Chem. 1983, 22, 2289–2292. 10.1021/ic00158a014. [DOI] [Google Scholar]; e Brunette J. P.; Lakkis Z.; Lakkis M.; Leroy M. J. F. Effect of inorganic aqueous anions on the synergic extraction of cadmium and zinc with 1-phenyl-3-methyl-4-benzoyl-pyrazol-5-one (HPMBP) and lipophilic ammonium salts. Polyhedron 1985, 4, 577–582. 10.1016/S0277-5387(00)86664-4. [DOI] [Google Scholar]; f Ma G.; Freiser H.; Muralidharan S. Centrifugal Partition Chromatographic Separation of Tervalent Lanthanides Using Acylpyrazolone Extractants. Anal. Chem. 1997, 69, 2835–2841. 10.1021/ac961192c. [DOI] [Google Scholar]; g Yamazaki S.; Hanada M.; Yanase Y.; Fukumori C.; Ogura K.; Saeki T.; Umetani S. A simple synthesis of novel extraction reagents. 4-Acyl-5-pyrazolone-substituted crown ethers. J. Chem. Soc., Perkin Trans. 1 1999, 693–696. 10.1039/A809589B. [DOI] [Google Scholar]; h Umetani S.; Kawase Y.; Le Q. T. H.; Matsui M. Acylpyrazolone derivatives of high selectivity for lanthanide metal ions: Effect of the distance between the two donating oxygens. J. Chem. Soc., Dalton Trans. 2000, 2787–2791. 10.1039/B000944J. [DOI] [Google Scholar]
  7. a Chiba P.; Holzer W.; Landau M.; Bechmann G.; Lorenz K.; Plagens B.; Hitzler M.; Richter E.; Ecker G. Substituted 4-Acylpyrazoles and 4-Acylpyrazolones: Synthesis and Multidrug Resistance-Modulating Activity. J. Med. Chem. 1998, 41, 4001–4011. 10.1021/jm980121y. [DOI] [PubMed] [Google Scholar]; b Zhao Z.; Dai X.; Li C.; Wang X.; Tian J.; Feng Y.; Xie J.; Ma C.; Nie Z.; Fan P.; Qian M.; He X.; Wu S.; Zhang Y.; Zheng X. Pyrazolone structural motif in medicinal chemistry: Retrospect and prospect. Eur. J. Med. Chem. 2020, 186, 111893 10.1016/j.ejmech.2019.111893. [DOI] [PMC free article] [PubMed] [Google Scholar]; c Sabnis R. W. Novel Compounds as PHD Inhibitors for Treating Heart, Lung, Liver, and Kidney Diseases. ACS Med. Chem. Lett. 2021, 12, 1868–1869. 10.1021/acsmedchemlett.1c00570. [DOI] [PMC free article] [PubMed] [Google Scholar]; d Mustafa G.; Zia-ur-Rehman M.; Sumrra S. H.; Ashfaq M.; Zafar W.; Ashfaq M. A critical review on recent trends on pharmacological applications of pyrazolone endowed derivatives. J. Mol. Struct. 2022, 1262, 133044 10.1016/j.molstruc.2022.133044. [DOI] [Google Scholar]
  8. a Caruso F.; Rossi M.; Tanski J.; Sartori R.; Sariego R.; Moya S.; Diez S.; Navarrete E.; Cingolani A.; Marchetti F.; Pettinari C. Synthesis, Structure, and Antitumor Activity of a Novel Tetranuclear Titanium Complex. J. Med. Chem. 2000, 43, 3665–3670. 10.1021/jm990539b. [DOI] [PubMed] [Google Scholar]; b Caruso F.; Pettinari C.; Marchetti F.; Natanti P.; Phillips C.; Tanski J.; Rossi M. Synthesis, Molecular Structure (X-ray and DFT), and Solution Behavior of Titanium 4-Acyl-5-pyrazolonates. Correlations with Related Antitumor β-Diketonato Derivatives. Inorg. Chem. 2007, 46, 7553–7560. 10.1021/ic700935x. [DOI] [PubMed] [Google Scholar]; c Budzisz E.; Malecka M.; Keppler B. K.; Arion V. B.; Andrijewski G.; Rozalski M. Synthesis, Structure, Protolytic Properties, Alkylating and Cytotoxic Activity of Novel Platinum(II) and Palladium(II) Complexes with Pyrazole-Derived Ligands. Eur. J. Inorg. Chem. 2007, 2007, 3728–3735. 10.1002/ejic.200700139. [DOI] [Google Scholar]; d Thaker B. T.; Surati K. R.; Modi C. K. Synthesis, spectral, thermal, and antibacterial investigation of mixed ligand complexes of oxovanadium(IV). Russ. J. Coord. Chem. 2008, 34, 25–33. 10.1134/S1070328408010053. [DOI] [Google Scholar]; e Caruso F.; Pettinari C.; Marchetti F.; Rossi M.; Opazo C.; Kumar S.; Balwani S.; Ghosh B. Inhibitory effect of β-diketones and their metal complexes on TNF-α induced expression of ICAM-1 on human endothelial cells. Bioorg. Med. Chem. 2009, 17, 6166–6172. 10.1016/j.bmc.2009.07.064. [DOI] [PubMed] [Google Scholar]; f Jadeja R. N.; Vyas K. M.; Gupta V. K.; Joshi R. G.; Prabha C. R. Syntheses, characterization and molecular structures of calcium(II) and copper(II) complexes bearing O2-chelate ligands: DNA binding, DNA cleavage and anti-microbial study. Polyhedron 2012, 31, 767–778. 10.1016/j.poly.2011.11.004. [DOI] [Google Scholar]; g De Pascali S. A.; Migoni D.; Monari M.; Pettinari C.; Marchetti F.; Muscella A.; Fanizzi F. P. Synthesis, Crystal Structure, and Biological Study of PtII Complexes with 4-Acyl-5-pyrazolones. Eur. J. Inorg. Chem. 2014, 2014, 1249–1259. 10.1002/ejic.201301479. [DOI] [Google Scholar]; h Pettinari R.; Pettinari C.; Marchetti F.; Skelton B. W.; White A. H.; Bonfili L.; Cuccioloni M.; Mozzicafreddo M.; Cecarini V.; Angeletti M.; Nabissi M.; Eleuteri A. M. Arene–Ruthenium(II) Acylpyrazolonato Complexes: Apoptosis-Promoting Effects on Human Cancer Cells. J. Med. Chem. 2014, 57, 4532–4542. 10.1021/jm500458c. [DOI] [PubMed] [Google Scholar]; i Caruso F.; Monti E.; Matthews J.; Rossi M.; Gariboldi M. B.; Pettinari C.; Pettinari R.; Marchetti F. Synthesis, Characterization, and Antitumor Activity of Water-Soluble (Arene)ruthenium(II) Derivatives of 1,3-Dimethyl-4-acylpyrazolon-5-ato Ligands. First Example of Ru(arene)(ligand) Antitumor Species Involving Simultaneous Ru–N7(guanine) Bonding and Ligand Intercalation to DNA. Inorg. Chem. 2014, 53, 3668–3677. 10.1021/ic403170y. [DOI] [PubMed] [Google Scholar]; j Marchetti F.; Di Nicola C.; Pettinari R.; Pettinari C.; Aiello I.; La Deda M.; Candreva A.; Morelli S.; De Bartolo L.; Crispini A. Zinc(II) Complexes of Acylpyrazolones Decorated with a Cyclohexyl Group Display Antiproliferative Activity Against Human Breast Cancer Cells. Eur. J. Inorg. Chem. 2020, 2020, 1027–1039. 10.1002/ejic.201900775. [DOI] [Google Scholar]
  9. Mies T.; Schürmann C.; Ito S.; White A. J. P.; Crimmin M. R.; Barrett A. G. M. Synthesis and Characterization of a Calcium-Pyrazolonato Complex. Observation of In-Situ Desolvation During Micro-Electron Diffraction. Z. Anorg. Allg. Chem. 2022, e202200294. [Google Scholar]
  10. a Al-Jallo H. N.; Al-Khashab A.; Sallomi I. G. Reactions of unsaturated tetra- and tri-esters with hydrazine hydrate and semicarbazide hydrochloride. J. Chem. Soc., Perkin Trans. 1 1972, 1022–1024. 10.1039/P19720001022. [DOI] [Google Scholar]; b Taylor A. W.; Cook R. T. A direct preparation of 2-aryl-4-ethoxy-carbonyl-3-pyrazolin-5-ones from aryl hydrazines. Tetrahedron 1987, 43, 607–616. 10.1016/S0040-4020(01)89994-4. [DOI] [Google Scholar]; c Eller G. A.; Holzer W. The 4-Methoxybenzyl (PMB) Function as a Versatile Protecting Group in the Synthesis of N-Unsubstituted Pyrazolones. Heterocycles 2004, 63, 2537–2555. 10.3987/COM-04-10190. [DOI] [Google Scholar]; d Guillou S.; Janin Y. L. 5-Iodo-3-Ethoxy-pyrazoles: An Entry Point to New Chemical Entities. Chem. – Eur. J. 2010, 16, 4669–4677. 10.1002/chem.200903442. [DOI] [PubMed] [Google Scholar]; e Tarabová D.; Šoralová S.; Breza M.; Fronc M.; Holzer W.; Milata V. Use of activated enol ethers in the synthesis of pyrazoles: reactions with hydrazine and a study of pyrazole tautomerism. Beilstein J. Org. Chem. 2014, 10, 752–760. 10.3762/bjoc.10.70. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. a Okafor E. C.; Nigeria N.; Uzoukwu B. A. Iron(III) Complexes of 4-Acyl Derivatives of 1-Phenyl-3-Methyl—Pyrazolone-5: Synthesis, Magnetic and Spectroscopic Studies. Synt. React. Inorg. Met. Org. Chem. 1992, 22, 921–927. 10.1080/15533179208016601. [DOI] [Google Scholar]; b Okafor E. C.; Adiukwu P. U.; Uzoukwu B. A. Sykoheses and Characterization of 4-iso-Butyryl and 4-iso-Valeroyl Derivatives of 1-FHE NTL-3-Methyl-Pyrazolone-5 and Their Uranium(VI), Tkorium(IV), Lanthanum(III), Iron(III), Lead(II) and Calciun(II) Complexes. Synt. React. Inorg. Met.-Org. Chem. 1993, 23, 97–111. 10.1080/15533179308016620. [DOI] [Google Scholar]; c Uzoukwu B. A. Spectroscopic studies of manganese(II) and zinc(II) complexes of 1-phenyl-3-methyl-4-acylpyrazolone-5. Spectrochim. Acta A 1993, 49, 281–282. 10.1016/0584-8539(93)80182-A. [DOI] [Google Scholar]
  12. a Shen X.-P.; Yuan A.-H. Bis(N,N-di-methyl-form-amide-κO)-bis-[1-phenyl-3-methyl-4-benzoyl-1H-pyrazol-5(4H)-onato-κ2O,O’]-nickel(II). Acta Cryst. 2004, 60, m1228–m1230. 10.1107/S160053680401877X. [DOI] [Google Scholar]; b Jing L.; Li J. Z.; Zhang H. Q.; Zhang Y.; Li J. Q. Synthesis, characterization and crystal structure of a Ni(II) complex derived from heterocyclic acylpyrazolone. J. Coord. Chem. 2009, 62, 2465–2471. 10.1080/00958970902842455. [DOI] [Google Scholar]; c Zhu H.; Wei Z.; Bu L.; Xu X.; Shi J. Bis(4-acetyl-3-methyl-1-phenyl-1H-pyrazol-5-olato-κ2O,O′)bis-(N,N-dimethyl-formamide-κO)nickel(II). Acta Cryst. 2010, 66, m904. 10.1107/S1600536810026231. [DOI] [PMC free article] [PubMed] [Google Scholar]; d Zhang X.; Huang M.; Du C.; Han J. Bis[4-(4-chloro-benzo-yl)-3-methyl-1-phenyl-1H-pyrazol-5-olato-κ2O,O′]bis--(methanol-κO)nickel(II). Acta Cryst. 2010, 66, m1361. 10.1107/S1600536810038973. [DOI] [PMC free article] [PubMed] [Google Scholar]; e Ding Y.-J.; Zhao C.-X.; Pei C.-Y.; Wen G.-X. Synthesis, Properties and Crystal Structure of a Novel Ni(II) Complex Derived from a 4-Heterocyclic Acylpyrazolone. Z. Naturforsch. B 2012, 67, 204–208. 10.1515/znb-2012-0304. [DOI] [Google Scholar]; f Esfahani M. H.; Behzad M.; Dusek M.; Kucerakova M. Synthesis and characterization of a series of acylpyrazolone transition metal complexes: Crystal structures and catalytic performance in the epoxidation of cyclooctene. Inorg. Chim. Acta 2020, 508, 119637 10.1016/j.ica.2020.119637. [DOI] [Google Scholar]
  13. Belousov Y. A.; Drozdov A. A. Lanthanide acylpyrazolonates: synthesis, properties and structural features. Russ. Chem. Rev. 2012, 81, 1159–1169. 10.1070/RC2012v081n12ABEH004255. [DOI] [Google Scholar]
  14. Platonova E. O.; Baranov E. V.; Bochkarev L. N.; Kurskii Y. A. The ruthenium pyrazolonate complex Ru(PMIP)2(PPh3)2: Synthesis, structure, and some properties. Russ. J. Coord. Chem. 2012, 38, 696–702. 10.1134/S1070328412110061. [DOI] [Google Scholar]
  15. Morosin B. The crystal structure of diaquobis(acetylacetonato)magnesium(II). Acta Cryst. 1967, 22, 315–320. 10.1107/S0365110X67000556. [DOI] [Google Scholar]
  16. a Uhlemann E.; Schilde U.; Weller F. Calcium- und Bariumkomplexe mit 1-Phenyl-3-methyl-4-benzoylpyrazol-5-on / Calcium and Barium Complexes with 1-Phenyl-3-methyl-4-benzoyl-5-pyrazolone. Z. Naturforsch. B 1995, 50, 31–36. 10.1515/znb-1995-0106. [DOI] [Google Scholar]; b Marchetti F.; Pettinari C.; Cingolani A.; Leonesi D.; Drozdov A.; Troyanov S. I. Copper and calcium complexes with the anionic O2-donor 4-tert-butylacetyl-3-methyl-1-phenylpyrazol-5-onato (Q). Influence of hydrogen-bond interactions on lattice architecture in the crystal structures of [CuQ2(H2O)] and [CaQ2(EtOH)2]. J. Chem. Soc., Dalton Trans. 1998, 3325–3334. 10.1039/A804768E. [DOI] [Google Scholar]
  17. Sahbari J. J.; Olmstead M. M. Structure of cis-bis(acetylacetonato)diaquacalcium monohydrate, [Ca(C5H7O2)2(H2O)2].H2O. Acta Cryst. 1983, 39, 208–211. 10.1107/S010827018300414X. [DOI] [Google Scholar]
  18. a Pettinari C.; Marchetti F.; Cingolani A.; Leonesi D.; Troyanov S. I.; Drozdov A. New co-ordination compounds derived from barium(II) and the anionic 4-tert-butylacetyl-3-methyl-1-phenylpyrazol-5-onate ligand (Q). Crystal and molecular structure of [Ba2Q4(H2O)4], [Ba2Q4(Him)4], [BaQ2(tetraglyme)] (tetraglyme = 2,5,8,11,14-pentaoxapentadecane) and [BaQ2(phen)2]. J. Chem. Soc., Dalton Trans. 1999, 1555–1562. 10.1039/A900463G. [DOI] [Google Scholar]; b Marchetti F.; Pettinari C.; Pettinari R.; Cingolani A.; Drozdov A.; Troyanov S. I. A new family of ionic dinuclear strontium (imH2)2[Sr2(Q)6] compounds (imH = imidazole; QH = 1-phenyl-3-methyl-4-acylpyrazol-5-one). J. Chem. Soc., Dalton Trans. 2002, 2616–2623. 10.1039/b200189f. [DOI] [Google Scholar]
  19. a Aghabozorg H.; Firoozi N.; Roshan L.; Eshtiagh-Hosseini H.; Salimi A. R.; Mirzaei M.; Ghanbari M.; Shamsipur M.; Ghadermazid M. Supramolecular structure of calcium(II) based on chelidamic acid: An agreement between theoretical and experimental studies. J. Iran. Chem. Soc. 2011, 8, 992–1005. 10.1007/BF03246555. [DOI] [Google Scholar]; b Starosta W.; Ptasiewicz-Bąk H.; Leciejewicz J. The Crystal and Molecular Structure of a New Calcium(II) Complex with Pyridine-2,6-DiCarboxylate, Water and Nitrate Ligands. J. Coord. Chem. 2002, 55, 1147–1153. 10.1080/0095897021000023130. [DOI] [Google Scholar]
  20. a Akama Y.; Kajitani M.; Sugiyama T.; Sugimori A. Crystal Structure of Scandium Complex of 1-Phenyl-3-methyl-4-benzoyl-5-pyrazolone. Anal. Sci. 1997, 13, 155–157. 10.2116/analsci.13.155. [DOI] [Google Scholar]; b Akama Y.; Sawada T.; Ueda T. Thermal and spectroscopic studies of scandium complex of 1-phenyl-3-methyl-4-benzoyl-5-pyrazolone. J. Mol. Struct. 2005, 750, 44–50. 10.1016/j.molstruc.2005.04.005. [DOI] [Google Scholar]; c Ueda T.; Sawada T.; Akama Y. Raman spectroscopic verification of fac-Sc(PMBP)3 by solid and solute spectra and density functional method calculation. J. Mol. Struct. 2005, 750, 51–57. 10.1016/j.molstruc.2005.04.006. [DOI] [Google Scholar]
  21. Anderson T. J.; Neuman M. A.; Melson G. A. Coordination chemistry of scandium. V. Crystal and molecular structure of tris (acetylacetonato)scandium(III). Inorg. Chem. 1973, 12, 927–930. 10.1021/ic50122a046. [DOI] [Google Scholar]
  22. Hansson E.; Oskarsson Å.; Rømming C.; Gronowitz S.; Koskikallio J.; Swahn C. G. Structural Studies on the Rare Earth Carboxylates. 20. The Crystal and Molecular Structure of Mono-Aquo Hydroxo Malonato Scandium(III) Monohydrate. Acta Chem. Scand. 1973, 27, 2841–2851. 10.3891/acta.chem.scand.27-2841. [DOI] [Google Scholar]
  23. Lorenz V.; Fischer A.; Jacob K.; Brüser W.; Edelmann F. T. f-Element Disiloxanediolates: Novel Si–O-based Inorganic Heterocycles. Chem. – Eur. J. 2001, 7, 848–857. . [DOI] [PubMed] [Google Scholar]
  24. Uzoukwu A. B.; Al-Juaid S. S.; Hitchcock P. B.; Smith J. D. The synthesis and crystal structures of 1-phenyl-3-methyl-4-butanolypyrazol-5-one and of two pyrazolonato complexes of iron. Polyhedron 1993, 12, 2719–2724. 10.1016/S0277-5387(00)80123-0. [DOI] [Google Scholar]
  25. a Tret’yakova I. N.; Chernii V. Y.; Tomachinskaya L. A.; Volkov S. V. Physicochemical properties of novel mixed-ligand complexes of zirconium and hafnium bis(4-benzoyl-3-methyl-1-phenyl-2-pyrazolin-5-onato)phthalocyaninates. Theor. Exp. Chem. 2006, 42, 175–180. 10.1007/s11237-006-0034-3. [DOI] [Google Scholar]; b Tomachynski L. A.; Tretyakova I. N.; Cherni V. Y.; Volkov S. V.; Kowalska M.; Legendziewicz J.; Gerasymchuk Y. S.; Radzki Synthesis and spectral properties of Zr(IV) and Hf(IV) phthalocyanines with β-diketonates as axial ligands. Inorg. Chim. Acta 2008, 361, 2569–2581. 10.1016/j.ica.2007.11.003. [DOI] [Google Scholar]
  26. a Silverton J. V.; Hoard J. L. Stereochemistry of Discrete Eight-Coördination. II. The Crystal and Molecular Structure of Zirconium(IV) Acetylacetonate. Inorg. Chem. 1963, 2, 243–249. 10.1021/ic50006a002. [DOI] [Google Scholar]; b Clegg W. Redetermination of the structure of tetrakis(acetylacetonato)zirconium(IV). Acta Cryst. 1987, 43, 789–791. 10.1107/S0108270187094083. [DOI] [Google Scholar]
  27. Saxena S.; Singh Y. P.; Rai A. K. Some bridged η1:η5-dicyclopentadienyltitanium pyrazolone complexes: Syntheses and structural elucidation. J. Organomet. Chem. 1984, 270, 301–309. 10.1016/0022-328X(84)80377-0. [DOI] [Google Scholar]
  28. Koh Y. B.; Christoph G. G. Metal-metal bonding in dirhodium tetracarboxylates. Structure of the bis(pyridine) adduct of tetra-μ-acetato-dirhodium(II). Inorg. Chem. 1978, 17, 2590–2596. 10.1021/ic50187a046. [DOI] [Google Scholar]
  29. a Yang L.; Jin W.; Lin J. Synthesis, crystal structure and magnetic properties of novel dinuclear complexes of manganese, cobalt and nickel with 4-acetylbispyrazolone. Polyhedron 2000, 19, 93–98. 10.1016/S0277-5387(99)00329-0. [DOI] [Google Scholar]; b Yang L. R.; Shao C. Y.; Wang Z. L.; Liu J. T.; Zhou L. F. Crystal Structure and Thermodecomposition Kinetics of Mn(II) Complex with 1-Phenyl-3-Methyl-4-Benzoyl-5-Pyrazolone. J. Chem. Crystallogr. 2010, 40, 58–63. 10.1007/s10870-009-9605-1. [DOI] [Google Scholar]
  30. a Zhang H.-Q.; Li J.-Z.; Gao S.; Zhang Y. Diethanolbis[4-(2-furo-yl)-3-methyl-1-phenyl-1H-pyrazol-5-onato]zinc(II). Acta Cryst. 2007, E63, m165–m166. 10.1107/S1600536806052718. [DOI] [Google Scholar]; b Pang X.-Z.; Li J.-Z.; Chen L.-S.; Zhang H.-Q. Synthesis, crystal structure, and electrocatalytic behavior of a zinc(II) complex derived from a heterocyclic acylpyrazolone. J. Coord. Chem. 2013, 66, 915–925. 10.1080/00958972.2013.771775. [DOI] [Google Scholar]; c Shaikh I. U.; Patel R. K.; Mevada V. A.; Gupta V. K.; Jadeja R. N. Binary and Ternary Zinc(II) Complexes of Acyl Pyrazolones: Synthesis, Spectroscopic Analysis, Crystal Structure and Antimalarial Activity. ChemistrySelect 2019, 4, 8286. 10.1002/slct.201901058. [DOI] [Google Scholar]; d Shaikh I.; Jadeja R. N.; Patel R. Three mixed ligand mononuclear Zn(II) complexes of 4-acyl pyrazolones: Synthesis, characterization, crystal study and anti-malarial activity. Polyhedron 2020, 183, 114528 10.1016/j.poly.2020.114528. [DOI] [Google Scholar]
  31. Marchetti F.; Pettinari C.; Pettinari R.; Arriva D.; Troyanov S.; Drozdov A. On the interaction of acylpyrazolonates with zinc(II) acceptors: the role of ancillary ligands. Inorg. Chim. Acta 2000, 307, 97–105. 10.1016/S0020-1693(00)00203-6. [DOI] [Google Scholar]
  32. Jadeja R. N.; Chhatrola M.; Gupta V. K. Zn(II) coordination compounds derived from 4-acyl pyrazolones and 1,10 phenanthroline: Syntheses, crystal structures, spectral analysis and DNA binding studies. Polyhedron 2013, 63, 117–126. 10.1016/j.poly.2013.07.018. [DOI] [Google Scholar]
  33. a Mickler W.; Reich A.; Sawusch S.; Schilde U.; Uhlemann E. Aquabis(3-methyl-4-octanoyl-1-phenyl-5-pyrazolonato-O,O’)zinc(II) and Bis(ethanol-O)bis(3-methyl-1-phenyl-4-stearoyl-5-pyrazolonato-O,O’)cadmium(II). Acta Cryst. 1998, 54, 776–779. 10.1107/S0108270197019215. [DOI] [Google Scholar]; b Dey S. K.; Bag B.; Dey D. K.; Gramlich V.; Li Y.; Mitra S. Synthesis and Characterization of Copper(II) and Zinc(II) Complexes Containing 1-Phenyl-3-methyl-4-benzoyl-5-pyrazolone. Z. Naturforsch. B. 2003, 58, 1009–1014. 10.1515/znb-2003-1012. [DOI] [Google Scholar]; c Xu L.-Y.; Li N.; Li J.-M.; Dong W. Synthesis, Structure, Antibacterial and Spectroscopic Properties of a Zinc(II) Complex with the Ligand 4-Heptanoyl-pyrazol-5-one. Z. Anorg. Allg. Chem. 2013, 639, 1800–1803. 10.1002/zaac.201200513. [DOI] [Google Scholar]
  34. a Goetz-Grandmont G. J.; Tayeb A.; Matt D.; Brunette J.-P.; Toupet L. Folding onto Copper(II) of a Flexible Bis-Chelating 4-Acyl-pyrazol-5-olate. Acta Cryst. 1995, 51, 53–57. 10.1107/S0108270194008024. [DOI] [Google Scholar]; b Pettinari C.; Marchetti F.; Santini C.; Pettinari R.; Drozdov A.; Troyanov S.; Battiston G. A.; Gerbasi R. Structure and volatility of copper complexes containing pyrazolyl-based ligands. Inorg. Chim. Acta 2001, 315, 88–95. 10.1016/S0020-1693(01)00330-9. [DOI] [Google Scholar]; c Shulgin V. F.; Pevzner N. S.; Konnik O. V.; Konshina O. I.; Aleksandrov G. G.; Eremenko I. L.; Bogomyakov A. S. Coordination compounds of some 3d metals with 3-methyl-1-phenyl-4-formylpyrazol-5-one. Russ. J. Inorg. Chem. 2014, 59, 58–62. 10.1134/S0036023614010161. [DOI] [Google Scholar]
  35. Nimmermark A.; Öhrström L.; Reedijk J. Metal-ligand bond lengths and strengths: are they correlated? A detailed CSD analysis. Z. Kristallogr. 2013, 228, 311–317. 10.1524/zkri.2013.1605. [DOI] [Google Scholar]
  36. a Safronova A. V.; Bochkarev L. N.; Malysheva I. P.; Baranov E. V. Facile synthesis of rare-earth pyrazolonates by the reaction of rare-earth metals with 1-phenyl-3-methyl-4-isobutyryl-5-pyrazolone. Crystal structures of [Ln(PMIP)3]2 (Ln = Y, Gd, Tb, Er, Tm). Inorg. Chim. Acta 2012, 392, 454–458. 10.1016/j.ica.2012.03.032. [DOI] [Google Scholar]; b Pettinari C.; Marchetti F.; Pettinari R.; Belousov Y. A.; Taydakov I. V.; Krasnobrov V. D.; Petukhov D. I.; Drozdov A. A. Synthesis of novel lanthanide acylpyrazolonato ligands with long aliphatic chains and immobilization of the Tb complex on the surface of silica pre-modified via hydrophobic interactions. Dalton Trans. 2015, 44, 14887–14895. 10.1039/C5DT01964H. [DOI] [PubMed] [Google Scholar]
  37. Shul’gin V. F.; Abkhairova S. V.; Konnik O. V.; Meshkov S. B.; Topilova Z. M.; Kiskin M. A.; Eremenko I. L. Synthesis, structure, and luminescent properties of lanthanide coordination compounds with 3-methyl-4-formyl-1-phenylpyrazol-5-one. Russ. J. Inorg. Chem. 2012, 57, 420–426. 10.1134/S0036023612030291. [DOI] [Google Scholar]
  38. Li J.; Zhang L.; Liu L.; Liu G.; Jia D.; Xu G. A series of pyrazolone lanthanide (III) complexes: Synthesis, crystal structures and fluorescence. Inorg. Chim. Acta 2007, 360, 1995–2001. 10.1016/j.ica.2006.10.010. [DOI] [Google Scholar]
  39. Pettinari C.; Marchetti F.; Pettinari R.; Drozdov A.; Troyanov S.; Voloshin A. I.; Shavaleev N. M. Synthesis, structure and luminescence properties of new rare earth metal complexes with 1-phenyl-3-methyl-4-acylpyrazol-5-ones. J. Chem. Soc., Dalton Trans. 2002, 1409–1415. 10.1039/B108058J. [DOI] [Google Scholar]
  40. Li X.-L.; Li J.; Zhu C.; Han B.; Liu Y.; Yin Z.; Li F.; Liu C.-M. An intense luminescent Dy(III) single-ion magnet with the acylpyrazolonate ligand showing two slow magnetic relaxation processes. New J. Chem. 2018, 42, 16992–16998. 10.1039/C8NJ03345E. [DOI] [Google Scholar]
  41. Li Z.-F.; Zhou L.; Yu J.-B.; Zhang H.-J.; Deng R.-P.; Peng Z.-P.; Guo Z.-Y. Synthesis, Structure, Photoluminescence, and Electroluminescence Properties of a New Dysprosium Complex. J. Phys. Chem. C 2007, 111, 2295–2300. 10.1021/jp064749t. [DOI] [Google Scholar]
  42. Bombiere G.; Polo A.; Wang J. F.; Wu J.; Xu G.-X. Synthesis and characterization of a ytterbium complex with diphenylacetylpyrazolone Yb(DPAP)3·(H2O)2·3EtOH. Inorg. Chim. Acta 1987, 132, 263–271. 10.1016/S0020-1693(00)81753-3. [DOI] [Google Scholar]
  43. a Zhang X.; Liu L.; Yu C.; Li H.; Fu G.; Lü X.; Wong W.-K.; Jones R. A. Highly efficient near-infrared (NIR) luminescent tris-β-diketonate Yb3 + complex in solution and in PMMA. Inorg. Chem. Commun. 2016, 70, 153–156. 10.1016/j.inoche.2016.06.006. [DOI] [Google Scholar]; b Yu C.; Zhang Z.; Liu L.; Li H.; He Y.; Lü X.; Wong W.-K.; Jones R. A. PMMA-supported hybrid materials doped with highly near-infrared (NIR) luminescent complexes [Zn(L1)(Py)Ln(L2)3] (Ln = Nd, Yb or Er). New J. Chem. 2015, 39, 3698–3707. 10.1039/C4NJ02373K. [DOI] [Google Scholar]; c Fu G.; Guan J.; Li B.; Liu L.; He Y.; Yu C.; Zhang Z.; Lü X. An efficient and weak efficiency-roll-off near-infrared (NIR) polymer light-emitting diode (PLED) based on a PVK-supported Zn2+–Yb3+-containing metallopolymer. J. Mater. Chem. C 2018, 6, 4114–4121. 10.1039/C7TC05400A. [DOI] [Google Scholar]
  44. a Pettinari C.; Marchetti F.; Cingolani A.; Drozdov A.; Troyanov S.; Timokhin I.; Vertlib V. Lanthanide metal complexes containing the first structurally characterized β-diketonate acid stabilized by hydrogen bonding. Inorg. Chem. Commun. 2003, 6, 48–51. 10.1016/S1387-7003(02)00678-0. [DOI] [Google Scholar]; b Pettinari C.; Marchetti F.; Pettinari R.; Drozdov A.; Semenov S.; Troyanov S. I.; Zolin V. A new rare-earth metal acylpyrazolonate containing the Zundel ion H5O2+ stabilized by strong hydrogen bonding. Inorg. Chem. Commun. 2006, 9, 634–637. 10.1016/j.inoche.2006.03.008. [DOI] [Google Scholar]; c Shul’gin V. F.; Konnik O. V.; Abkhairova S. V.; Gusev A. N.; Meshkova S. B.; Kiriyak A. V.; Rusanov E. B.; Hasegawa M.; Linert W. Anionic lanthanide complexes with 3-methyl-1-phenyl-4-formylpyrazole-5-one and hydroxonium as counter ion. Inorg. Chim. Acta 2013, 402, 33–38. 10.1016/j.ica.2013.03.044. [DOI] [PMC free article] [PubMed] [Google Scholar]; d Belousov Y. A.; Korshunov V. M.; Metlin M. T.; Metlina D. A.; Kiskin M. A.; Aminev D. F.; Datskevich N. P.; Drozdov A. A.; Pettinari C.; Marchetti F.; Taydakov I. V. Dyes Pigm. 2022, 199, 110078. 10.1016/j.dyepig.2021.110078. [DOI] [Google Scholar]
  45. Shul’gin V. F.; Abkhairova S. V.; Konnik O. V.; Meshkova S. B.; Topilova Z. M.; Rusanov E. B.; Aleksandrov G. G.; Eremenko I. L. Anionic lanthanide complexes with 3-methyl-4-formyl-1-phenyl-5-pyrazolone. Russ. J. Inorg. Chem. 2013, 58, 678–683. 10.1134/S0036023613060223. [DOI] [Google Scholar]
  46. Deun R. V.; Nockemann P.; Parac-Vogt T. N.; Van Hecke K.; Van Meervelt L.; Görller-Walrand C.; Binnemans K. Near-infrared photoluminescence of lanthanide complexes containing the hemicyanine chromophore. Polyhedron 2007, 26, 5441–5447. 10.1016/j.poly.2007.08.005. [DOI] [Google Scholar]
  47. a Huang C.; Wang K.; Xu G.; Zhao X.; Xie X.; Xu Y.; Liu Y.; Xu L.; Li T. Langmuir Film-Forming and Second Harmonic Generation Properties of Lanthanide Complexes. J. Phys. Chem. 1995, 99, 14397–14402. 10.1021/j100039a029. [DOI] [Google Scholar]; b Pettinari C.; Marchetti F.; Cingolani A.; Drozdov A.; Timokhin I.; Troyanov S. I.; Tsaryuk V.; Zolin V. Syntheses, structural and spectroscopic investigation (IR, NMR and luminescence) of new terbium and europium acylpyrazolonates. Inorg. Chim. Acta 2004, 357, 4181–4190. 10.1016/j.ica.2004.06.024. [DOI] [Google Scholar]
  48. a Wang K. Z.; Huang C. H.; Xu G. X.; Xu Y.; Liu Y. Q.; Zhu D. B.; Zhao X. S.; Xie X. M.; Wu N. Z. Preparation, Characterization, and Second-Harmonic Generation of a Langmuir-Blodgett Film Based on a Rare-Earth Coordination Compound. Chem. Mater. 1994, 6, 1986–1989. 10.1021/cm00047a017. [DOI] [Google Scholar]; b Li H.; Huang C.-H.; Zhao X.-S.; Xie X.-M.; Xu L.-G.; Li T.-K. Molecular Design and Langmuir-Blodgett Film Studies on a Series of New Nonlinear Optical Rare Earth Complexes. Langmuir 1994, 10, 3794–3796. 10.1021/la00022a065. [DOI] [Google Scholar]; c Zhao X.; Xing J.; Li P.; Xie X.; Xia X.; Li H.; Huang C.; Li T.; Xu L. Effects of Adsorption and Reaction on the Second Harmonic Generation of Langmuir-Blodgett Films. Langmuir 1995, 11, 3620–3622. 10.1021/la00010a004. [DOI] [Google Scholar]; d Li H.; Huang C.-H.; Zhao Y.-L.; Li T.-K.; Bai J.; Zhao X.-S.; Xia X.-H. Langmuir-Blodgett film and second harmonic generation of a series of new nonlinear optical rare earth complexes. Solid State Commun. 1995, 94, 731–733. 10.1016/0038-1098(95)00157-3. [DOI] [Google Scholar]; e Wang K. Z.; Huang C. H.; Xu G. X.; Zhou Q. F. Liquid-crystalline behaviors of lanthanide complexes containing hemicyanine. Solid State Commun. 1995, 95, 223–225. 10.1016/0038-1098(95)00221-9. [DOI] [Google Scholar]; f Pavier M. A.; Weaver M. S.; Lidzey D.; Richardson T.; Searle T. M.; Bradley D. D. C.; Huang C. H.; Li H.; Zhou D. Electroluminescence from dysprosium- and neodymium-containing LB films. Thin Solid Films 1996, 284-285, 644–647. 10.1016/S0040-6090(95)08411-8. [DOI] [Google Scholar]; g Xia W. S.; Huang C. H.; Gan L. B.; Li H.; Zhao X. S. Photoelectric response of a bilayer lipid membrane doped with an azo pyridinium compound containing a rare-earth-metal complex. J. Chem. Soc., Faraday Trans. 1996, 92, 769–772. 10.1039/FT9969200769. [DOI] [Google Scholar]; h Binnemans K. Stilbazolium dyes containing rare-earth ions. J. Alloys Compd. 2000, 303-304, 125–131. 10.1016/S0925-8388(00)00618-6. [DOI] [Google Scholar]
  49. a Ryan R. R.; Jarvin G. D. Structure of bis(4-benzoyl-2,4-dihydro-5-methyl-2-phenyl-3H-pyrazol-3-onato-O,O’)(dimethyl sulfoxide-O)dioxouranium(VI). Acta Cryst. 1987, 43, 1296–1298. 10.1107/S0108270187092114. [DOI] [Google Scholar]; b Okafor E. C.; Uzoukwu A. B.; Hitchcock P. B.; Smith J. D. Pyrazolonato complexes of uranium. Crystal structures of bis-oxobis(1-phenyl-3-methyl-4-acetylpyrazol-5-onato)aquouranium(VI) and Bis-oxobis(1-phenyl-3-methyl-4-benzoylpyrazol-5-onato)(propanol)uranium(VI). Inorg. Chim. Acta 1990, 172, 97–103. 10.1016/S0020-1693(00)80456-9. [DOI] [Google Scholar]; c Zhu L.-M.; Wang S.-W.; Li B.-L.; Jin J.-R.; Zhang Y. Syntheses and structures of two new uranyl complexes [UO2(DPDPU)2(NO3)2](C6H5CH3) and [UO2(PMBP)2(DPDPU)](CH3C6H4CH3)0.5. J. Coord. Chem. 2006, 59, 1609–1614. 10.1080/00958970600577494. [DOI] [Google Scholar]; d Zhu L.-M.; Wang L.-Y.; Jin J.-R.; Li B.-L.; Zhang Y. Synthesis and structures of two uranyl β-diketonate complexes [UO2(DBM)2(DEDPU)] and [UO2(PMBP)2(DEDPU)]. J. Coord. Chem. 2008, 61, 917–925. 10.1080/00958970701413243. [DOI] [Google Scholar]; e Zhu L.; Yuan D.; Li B.-L.; Li H. Syntheses and structures of three disulfoxide uranyl complexes. J. Coord. Chem. 2010, 63, 3006–3015. 10.1080/00958972.2010.505282. [DOI] [Google Scholar]; f Chumakov Y. M.; Petrenko P. A.; Julea F. G.; Tsapkov V. I.; Gulea A. P. Crystal structure of bis(4-benzoyl-3-methyl-1-phenyl-4,5-dihydro-1H-pyrazol-5-onato)-(ethanol)-dioxourane(VI) ethanol solvate. J. Struct. Chem. 2014, 55, 1116–1119. 10.1134/S0022476614060183. [DOI] [Google Scholar]
  50. Mies T.; White A. J. P.; Rzepa H. S.; Barluzzi L.; Layfield R. A.; Barrett A. G. M.. Imperial College Research Data Repository. 2022, 10.14469/hpc/10386. [DOI]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

ic3c01506_si_001.pdf (44MB, pdf)
ic3c01506_si_003.pdf (9.5MB, pdf)

Articles from Inorganic Chemistry are provided here courtesy of American Chemical Society

RESOURCES