Next Article in Journal
Ascorbic Acid Sensor Based on CdS QDs@PDA Fluorescence Resonance Energy Transfer
Next Article in Special Issue
Synthesis, Crystal Structure and Magnetic Properties of a Trinuclear Copper(II) Complex Based on P-Cresol-Substituted Bis(α-Nitronyl Nitroxide) Biradical
Previous Article in Journal
Carnosic Acid Protects INS-1 β-Cells against Streptozotocin-Induced Damage by Inhibiting Apoptosis and Improving Insulin Secretion and Glucose Uptake
Previous Article in Special Issue
A New Pt(II) Complex with Anionic s-Triazine Based NNO-Donor Ligand: Synthesis, X-ray Structure, Hirshfeld Analysis and DFT Studies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Selective Formation of Intramolecular Hydrogen-Bonding Palladium(II) Complexes with Nucleosides Using Unsymmetrical Tridentate Ligands

1
Institute of Liberal Arts and Science, Kanazawa University, Kakuma, Kanazawa 920-1192, Ishikawa, Japan
2
Department of Chemistry, Kanazawa University, Kakuma, Kanazawa 920-1192, Ishikawa, Japan
3
Research Institute for Interdisciplinary Science, Okayama University, 3-1-1 Tsushima-naka, Okayama 700-8530, Japan
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(7), 2098; https://doi.org/10.3390/molecules27072098
Submission received: 25 February 2022 / Revised: 17 March 2022 / Accepted: 21 March 2022 / Published: 24 March 2022
(This article belongs to the Special Issue Crystal Structures of Metal Complexes)

Abstract

:
Three palladium(II) complexes with amino-amidato-phenolato-type tridentate ligands were synthesized and characterized by 1H NMR spectroscopy and X-ray crystallography. The strategic arrangement of a hydrogen-bond donor and acceptor adjacent to the substitution site of the PdII complex allowed the selective coordination of nucleosides. Among two pyrimidine-nucleosides, cytidine and 5-methyluridine, cytidine was successfully coordinated to the PdII complex while 5-methyluridne was not. On the other hand, both purine-nucleosides, adenosine and guanosine, were coordinated to the PdII complex. As purines have several coordination sites, adenosine afforded three kinds of coordination isomers expected from the three different donors. However, guanosine afforded a sole product according to the ligand design such that the formation of double intramolecular hydrogen-bond strongly induced the specific coordination by N1-position of guanine moiety. Furthermore, the preference of the nucleosides was evaluated by scrambling reactions. It was found that the preference of guanosine is nearly twice as high as adenosine and cytidine, owing to the three-point interaction of a coordination bond and two hydrogen bonds. These results show that the combination of a coordination and hydrogen bonds, which is reminiscent of the Watson–Crick base pairing, is an effective tool for the precise recognition of nucleosides.

Graphical Abstract

1. Introduction

Hydrogen-bonding interactions have attracted much attention in the past and present as they are involved in many chemical and biological systems. Hydrogen bonds exhibit a directionality and reversibility upon bond formation. Although the bond energy is much smaller than that of a covalent bond, such unique properties enable construction of supramolecular structures of a molecule to control the chemical and physical properties [1,2,3,4,5,6,7,8]. Furthermore, it is also possible to precisely recognize a molecule by tuning the relative position of hydrogen-bond donors and acceptors in a molecule as represented by the formation of the base pair in DNA [9,10].
Coordination of biomolecules such as nucleobases to a transition metal ion often exhibits cytotoxic properties [11,12,13,14]. For example, a well-known mononuclear PtII complex, cisplatin cis-[PtCl2(NH3)2], inhibits the replication of DNA by selective coordination by two continuous guanine residues of DNA. On the other hand, we previously reported synthesis and crystal structures of cobalt and manganese complexes with amino-amidato-phenolato-type tridentate ligands (Figure 1) [15,16]. In combination with a PdII ion, for which square planar geometry is expected, the amino-amidato-phenolato ligand affords a vacant coordination site for the recognition of nucleosides by arranging a hydrogen-bond donor and acceptor to accommodate a specific nucleoside. For this PdII complex fragment, selective coordination of a nucleobase is expected, owing to a labile coordination bond in PdII complex and the arrangement of hydrogen-bonding sites. In this study, we synthesized PdII complexes with two amino-amidato-phenolato-type ligands and evaluated the effects of hydrogen-bonding interaction on the selectivity of nucleosides upon coordination to the PdII complex fragment. For such tridentate ligands, it is expected to form a Pd–N coordination and two intramolecular hydrogen bonds with guanine and cytosine moieties. We demonstrated that this three-point interaction enables the precise recognition of nucleobases.

2. Results and Discussion

2.1. Preparation

Preparation of PdII Complexes

The ligand precursors H2Amp, H2Apr, and PdII starting material were prepared according to the previously reported procedures [15,17]. The PdII complex was synthesized by a reaction of [PdCl2(CH3CN)2] and H2Amp followed by an addition of triethylamine and N(CH3)4OH·5H2O as a base in methanol. The product was crystallized by diffusing diethyl ether vapor into the methanol solution to afford yellow crystals of N(CH3)4[Pd(Amp)Cl]·H2O (1). To accept the coordination by nucleosides, the presence of chloride ligand is not suitable because of its coordination ability. When the 1H NMR spectrum of this product was measured in CD3OD, it exhibited rather broad resonances (Figure 2a). This implies that the complex [Pd(Amp)Cl] undergoes a rapid exchange reaction of Cl ligand with a solvent methanol molecule. Thus, we exploited this reaction to isolate a chloride-free complex, [Pd(Amp)(CH3CN)]·CH3CN (1′), by using KOtBu instead of N(CH3)4OH·5H2O. In this reaction, potassium salt of 1 was not obtained; instead, 1′ was selectively produced due to the elimination of KCl from the reaction mixture. 1′ is more suitable for the substitution reaction with a nucleoside because acetonitrile is a better leaving group.
An analogous PdII complex with Apr2− was also prepared by a similar procedure with 1′, and [Pd(Apr)(CH3CN)] (2) was obtained. Although the ligand structure of Apr2− is analogous to Amp2−, 2 exhibited a different coordination mode of the tridentate ligands. The 1H NMR spectrum in CD3CN only showed three aryl-H resonances, which indicates that one of the protons on the aromatic group was removed by the base (Figure 2c). The elimination of the proton induced the formation of a Pd–C bond, and the resulting direct coordination of the ring carbon makes the acetonitrile coordination site rather hydrophobic. The direct coordination of the aromatic ring was also supported by the significant up-field shift of aryl-H resonances. The formation of 2 has parallels with orthometalation reaction in organometallic chemistry.

2.2. Crystal Strucre of PdII Complexes

2.2.1. Crystal Structures of 1 and 1′

Single-crystals of 1 and 1′ were prepared by diffusing diethyl ether vapor to the reaction solution or slow evaporation of the solution, respectively. The crystallographic information is summarized in Table 1. In both crystals, the PdII ion takes general square planar coordination geometry (Figure 3). In both 1 and 1′, Amp2− ligand coordinates to a PdII center with a phenolato-O, amidato-N, and amino-N atoms as a tridentate ligand. In 1, the remaining coordination site is occupied by a Cl ion to afford an anionic PdII complex, [Pd(Amp)Cl]. In 1′, on the other hand, the PdII center is coordinated by an acetonitrile molecule instead of Cl ligand, although the reaction solvent was acetonitrile in both cases. This is because the Cl ion was removed from the system during the reaction due to low solubility of KCl. The bonding parameters around the PdII ion are summarized in Table 2. The bond lengths around the PdII center are nearly identical between 1 and 1′ except those of the ancillary ligand Cl and CH3CN.
In 1, intermolecular hydrogen bonds are formed between neighboring complex anions and water molecules of crystallization to construct two-dimensional sheet structures along the bc plane. The two-dimensional sheet structures in the bc plane are also formed in 1′ by the intermolecular hydrogen bonds between the neighboring complexes. These results are indicative of the strong ability of Amp2− to induce hydrogen-bonding interactions.

2.2.2. Crystal Structure of 2

In the crystal of 2, [Pd(Apr)(CH3CN)] also takes a general square planar coordination geometry (Figure 4 and Table 3). The asymmetric unit consists of two independent [Pd(Apr)(CH3CN)] molecules, and each molecule is stacked to form a centrosymmetric dimer structure by NH π interactions (the symmetric operation for Pd1 moiety is 1−x, 1−y, 2−z and for Pd2 moiety is 2−x, 2−y, 1−z). Despite the structural similarity with Amp2−, Apr2− ligand coordinates to a PdII center with phenyl-C, amidato-N, and amino-N atoms. The difference in the coordination modes can be attributed to the combination of chelate rings. For a linear tridentate ligand in octahedral complexes, the combination of six-membered and five-membered chelates is more favored to afford meridional coordination geometry, whereas the combination of two six-membered chelates favors facial coordination geometry [15,16]. The coordination of a linear-tridentate ligand in square planar complexes may correspond to the coordination mode of the meridional type in the octahedral complexes. The observation of aromatic carbon coordination in the formation of complex 2 coincides with the larger stability gain by the combination of five- and six-membered chelates in the octahedral complexes. When Apr2− ligand coordinates in the same way to Amp2−, the resulting formation of two six-membered rings corresponds to facial coordination in the octahedral complexes, but in square planar complexes, it is impossible. The average bond length between the PdII ion and the terminal amino-N atom, 2.13 Å, is significantly longer than those in 1 and 1′ (2.03 Å) because of a strong trans influence of phenyl-C donor [18]. The hydroxy group of the ligand forms an intramolecular hydrogen bond with the O atom of the carbonyl group.

2.3. Reaction of 1 with Nucleosides

Although 1′ is suitable for the ligand exchange, KCl was easily contaminated as impurity during the synthesis. Thus, we employed 1 for the substitution study with nucleosides as the substitution reaction of the chloride ligand is readily proceeded in methanol. Reactivity and selectivity of [Pd(Amp)Cl] with four nucleosides, guanosine, adenosine, cytidine, and 5-methyluridine, were evaluated in methanol by 1H NMR spectroscopy (Figure 5). The reaction was performed by mixing 1 and the nucleoside in methanol.

2.3.1. Reaction of 1 with Pyrimidine-Nucleosides

Two pyrimidine nucleosides, cytidine and 5-methyluridine, were evaluated. Cytidine was suitable for selective coordination to [Pd(Amp)] fragment because it has only one possible coordination site and both hydrogen-bond donor and acceptor sites were available within the proximity of the coordination site (2- and 4-position of the pyrimidine ring). The 1H NMR spectrum indicated the formation of [Pd(Amp)(cytidine)]-adduct (Pd-C) as shown in Figure 6. The broad resonances observed in the spectrum of 1 became converged in clear multiplets in the presence of cytidine, implying that Pd-C is reasonably stable without further ligand exchange reaction. The splitting and geminal coupling of the signal corresponding to the coordinating amino group implies the presence of nonequivalent protons by the intramolecular hydrogen bond (4.6–4.3 ppm in Figure S4, Supplementary Materials). In the spectrum, only a trace amount of minor product was observed (<5%), which was attributed to a negligible coordination mode such as carbonyl-O coordination. The observation of a selective formation of Pd-C indicated that it has a specific binding mode by the coordination of N-donor site on pyrimidine that was supported by two pairs of donor-acceptor interactions (Figure 7). This three-point interaction is reminiscent of the Watson–Crick base pairing in DNA.
In the case of 5-methyluridine, on the other hand, no shift was observed for the resonances of both the nucleoside and [Pd(Amp)Cl], indicating that 5-methyluridine was not able to coordinate to the Pd center (Figure 8). In the presence of 1 equiv. of triethylamine, a few sets of resonances corresponding to 5-methyluridine, including unreacted, were observed. From the integration ratio, over 80% of 5-methyluridine remained unreacted, although the formation of a minor amount of [Pd(Amp)(5-methyluridine)]-adduct was observed. The resonances corresponding to [Pd(Amp)]-fragment became sharp and the formation of a new species [Pd(Amp)(OCD3)] was suggested. These results showed that two hydrogen-bond acceptors in 5-methyluridine distract its coordination to the [Pd(Amp)]-fragment due to the electrostatic repulsive force between O atoms of Amp ligand and 5-methyluridine. Thus, [Pd(Amp)]-fragment exhibited high selectivity on the coordination of pyrimidine-nucleosides, owing to the preference of hydrogen-bonding interactions between Amp and the pyrimidine moieties.

2.3.2. Reaction of 1 with Purine-Nucleosides

Two purine-nucleosides, adenosine and guanosine, were evaluated as well as the pyrimidine-nucleosides. Due to the increasing possible coordination sites (N1-, N3-, and N7-position), purine-nucleosides might afford different products compared with the previous cases. Adenosine possesses a hydrogen-bond donor at 6-position and no hydrogen-bond acceptor exists at the coordination site. For [Pd(Amp)(adenosine)]-adduct (Pd-A), three different coordination modes are possible depending on which N atom is coordinated (Figure 9). The 1H NMR spectrum indicated that the product is a mixture of Pd-A (Figure 10). The resonances according to the protons at adenine moiety imply that three species (Pd-A1, Pd-A2, and Pd-A3) are present in the reaction solution. The formation ratio of the products was determined to be Pd-A1:Pd-A2:Pd-A3 = 6:3:2 from the integration of the resonances of the aryl-H of the adenine moiety. The N1-coordination mode is favorable because it is supported by a hydrogen bond (Figure 9a). Such intramolecular hydrogen-bond formation with adenine was reported in a PdII complex with coordination at the N9 position of adenine, although the N9 position is not available for coordination in our current system [19]. In contrast, no intramolecular hydrogen-bond interactions are expected for the other two modes (Figure 9b,c). Moreover, the N3 coordination must endure a large degree of steric hindrance between the ligand and ribose moiety. Thus, the major product Pd-A1 is in the N1-coordination mode, and the two other sets of signals Pd-A2 and Pd-A3 correspond to the N7- and N3-coordination modes, respectively.
Guanosine is structurally similar to cytidine in terms of the position of hydrogen-bond donors and acceptors (2- and 6-position, respectively). Although guanosine need to be deprotonated for the coordination at the N1 atom, formation of double intermolecular hydrogen bonds is expected for guanosine (Figure 11). The 1H NMR spectrum of 1:1 mixture of 1 and guanosine in CD3OD exhibited a sole set of resonances corresponding to [Pd(Amp)(guanosine)]-adduct (Pd-G), as shown in Figure 12. As observed in the case of cytidine, geminal coupling of the signals corresponding to the coordinating amino group was also observed, indicating N1 coordination of guanine moiety (5.3–5.0 ppm in Figure S7). It is to be noted that N7 is the most nucleophilic position in a neutral guanine moiety, and as a result, it is the primary binding site [20]. The N1 coordination in Pd-G without addition of base is indicative of the strong preference toward formation of intramolecular hydrogen bonds.

2.3.3. Scrambling Reaction of 1 with Two Nucleosides

To investigate the selectivity among nucleosides, complex 1 was reacted with two kinds of nucleosides at the same time. The Pd complex 1, nucleoside-1, and nucleoside-2 were reacted in methanol at room temperature in a 1:1:1 molar ratio. 5-Methyluridine was excluded because it did not react with 1. All sets of the combinations from two kinds of nucleosides were tested: adenosine vs. cytidine, adenosine vs. guanosine, and guanosine vs. cytidine (Figure 13).
In all cases, a mixture of a few Pd-nucleoside adducts was obtained. In the presence of adenosine and cytidine, Pd-A1, Pd-A2, and Pd-C were formed and formation of Pd-A3 was negligible. From the formation ratio of Pd-A1:Pd-A2:Pd-C = 3:2:6, the coordination of cytidine is significantly more favorable than adenosine because of the formation of the double intramolecular hydrogen bonds. In the case of adenosine and guanosine, both Pd-A1 and Pd-A2 were observed, similar to the previous case. The formation ratio of Pd-A1:Pd-A2:Pd-G = 3:2:10 indicated that the coordination of guanosine is about twice as favorable as that of adenosine. These two experiments are indicative of the stabilization effect of double intramolecular hydrogen-bond formation on the coordination.
Finally, the coordination ability of guanosine and cytidine was compared. The NMR spectrum showed only two species, Pd-G and Pd-C, in the reaction. The formation ratio was found to be Pd-G:Pd-C = 2:1, which is reasonably consistent with the ratio expected from the previous two experiments, (Pd–G/Pd-A):(Pd-C/Pd-A) = 10:6. Thus, guanosine has the highest preference among the four nucleosides. Considering the number of hydrogen-bonding interactions of guanosine and cytidine with [Pd(Amp)]-moiety, the formation ratio should be comparable. As the anionic N1 position of the guanine moiety is more nucleophilic, however, the coordination of guanosine was more favorable than that of cytidine.

3. Materials and Methods

3.1. Measurements

Elemental analyses (C, H, and N) were performed at the Research Institute for Instrumental Analysis, Kanazawa University. 1H NMR measurements were carried out at 22 °C on a JEOL 400SS spectrometer. Chemical shifts were referenced to the solvent residual peak [21].

3.2. Materials

All the chemicals were used as received without further purification. The ligand precursors N-(2-amino-2-methylpropyl)salicylamide (H2Amp), N-3-aminopropylsalicylamide (H2Apr), and the starting material of Pd complex, [PdCl2(CH3CN)2], were synthesized according to the previously reported procedure [15,17].

3.3. Preparations

N(CH3)4[Pd(Amp)Cl]·H2O (1). To an acetonitrile solution (20 mL) of [PdCl2(CH3CN)2] (51.4 mg, 0.20 mmol) was added a solid H2Amp (41.5 mg, 0.20 mmol) and triethylamine (28 µL). After stirring the mixture for 10 min, a solid (CH3)4NOH·5H2O (36.2 mg, 0.20 mmol) was added followed by stirring at room temperature overnight. The reaction solution was concentrated to 2 mL by Ar gas bubbling. Diethyl ether solution was diffused to the solution to give orange, needle-shaped crystals. Yield: 65.8 mg, 53%. Anal. Calcd for N(CH3)4[Pd(Amp)Cl]·H2O = C15H28ClN3O3Pd: C, 40.92; H, 6.41; N, 9.54%. Found: C, 40.64; H, 6.25; N, 9.54%. 1H NMR (399 MHz, CD3OD): δ 7.97 (broad, 1H, aryl-H), 7.05 (broad, 1H, aryl-H), 6.72 (broad, 1H, aryl-H), 6.58 (broad, 1H, aryl-H), 4.62 (s, 2H, NH2), 3.41 (s, 2H, CH2), 3.19 (s, 12H, CH3), 1.40 (s, 6H, CH3).
[Pd(Amp)(CH3CN)]·CH3CN (1′). To an acetonitrile solution (10 mL) of [PdCl2(CH3CN)2] (25.9 mg, 0.10 mmol) was added a solid H2Amp (20.8 mg, 0.1 mmol). After dissolving the ligand completely, a methanolic solution (5 mL) of KOtBu (22.3 mg, 0.20 mol) was added to the mixture, followed by stirring overnight. The reaction solution was concentrated to 3 mL by Ar gas bubbling. The residue was filtered, and the filtrate was crystallized by slow evaporation of the solvent. Yield: 22.4 mg, 57%. Anal. Calcd for [Pd(Amp)(CH3CN)]·1.3H2O = C13H19.6N3O3.3Pd: C, 41.40; H, 5.24; N, 11.14%. Found: C, 41.85; H, 5.49; N, 10.60%. 1H NMR (399 MHz, CD3OD): δ 8.00 (m, 1H, aryl-H), 7.06 (dd, J = 8.6, 6.8 Hz, 1H), 6.71 (d, J = 8.2 Hz, 1H), 6.58 (dd, J = 8.3, 6.8 Hz, 1H), 3.82 (s, 2H, CH2), 2.52 (s, 3H, CH3), 1.43 (d, J = 7.6 Hz, 6H).
[Pd(Apr)(CH3CN)] (2). To an acetonitrile solution (8 mL) of [PdCl2(CH3CN)2] (25.6 mg, 0.10 mmol) was added a solid H2Apr (19.4 mg, 0.1 mmol) and triethylamine (14 µL). After dissolving the ligand completely, a methanolic solution (3 mL) of KOtBu (11.0 mg, 0.10 mol) was added to the mixture, followed by stirring overnight. The reaction solution was filtered to remove white precipitate. The filtrate was evaporated to dryness and dissolved in acetonitrile (2 mL). Diethyl ether vapor was diffused to the solution to give yellow microcrystals. Yield: 7.3 mg, 23%. This compound was extremely hygroscopic and unable to perform elemental analysis. 1H NMR (399 MHz, CD3CN): δ 6.76 (m, 1H, aryl-H), 6.59 (dd, J = 7.4, 0.7 Hz, 1H, aryl-H), 6.33 (dd, J = 8.2, 0.9 Hz, 1H, aryl-H), 3.22 (m, 2H, CH2), 2.66 (m, 4H, CH2, and NH2), 1.57 (m, 2H, CH2).

3.4. Crystallography

Crystallographic data are summarized in Table 4. Single-crystal X-ray diffraction data were obtained with a Rigaku XtaLAB AFC11 diffractometer with graphite-monochromated Mo Kα radiation (λ = 0.71073 Å). A single crystal was mounted with a glass capillary and flash-cooled with a cold N2 gas stream. Data were processed using the CrysAlisPro software packages. The structure was solved by intrinsic phasing methods using the SHELXT [22] software packages and refined on F2 (with all independent reflections) using the SHELXL [23] software packages. The non-hydrogen atoms were refined anisotropically. The hydrogen atoms except for OH and CH3 groups were located at the calculated positions and refined isotopically using the riding models. The hydrogen atoms for OH and CH3 groups were located by difference Fourier maps and allowed to rotate. The O–H atoms for water molecules were located at the position suitable for hydrogen bonds and refined freely with a restrained distance. In the case of 2, only tiny single-crystals were obtained. As it is difficult to apply a sufficient absorption correction to a tiny crystal based on the actual size, the Rint value for 2 is apparently large. However, the quality of the crystal itself was reasonable. The anisotropic refinement for the acetonitrile ligands was restrained to obtain reasonable parameters. The Cambridge Crystallographic Data Centre (CCDC) deposition numbers are included in Table 4.

3.5. Reaction of 1 with Nucleosides

The reactions were performed by mixing a methanol solution of 1 with a solid nucleoside. Although some nucleosides are hardly soluble to typical organic solvents such as methanol and dimethylformamide, they dissolved well in methanol in the presence of 1. A typical procedure for reactions of 1 with nucleosides is as follows.
A 5 mL methanolic solution of 1 (20 µmol) was added to a solid nucleoside (20 µmol) followed by stirring overnight. Half of the reaction mixture was taken and evaporated to dryness by Ar gas bubbling. The resulting yellow residue was dried in vacuo for 1 h. The residue was dissolved in ca. 0.7 mL of CD3OD, followed by 1H NMR measurement.

4. Conclusions

Palladium(II) complexes containing unsymmetric tridentate ligand, Amp2−, and Apr2−, were synthesized and crystallographically characterized. Although Amp2− exhibited a typical O–N–N coordination mode, Apr2− formed a Pd–C bond, resulting an observation of the C–N–N coordination mode. The Pd complex with Amp2−, 1, was employed for evaluating the effect of stereoselective intramolecular hydrogen-bonding interaction of four nucleosides (adenosine, guanosine, cytidine, and 5-methyluridine). Three of the four nucleosides resulted in coordination to the [Pd(Amp)]-fragment. In the case of 5-methyluridine, in which no intramolecular hydrogen bond is expected upon coordination, [Pd(Amp)(5-methyluridine)]-adduct was not formed. By contrast, the other three nucleosides, which form intramolecular hydrogen bonds upon coordination, were coordinated to the [Pd(Amp)]-fragment. These results clearly show that hydrogen-bonding interactions play an important role upon coordination to the [Pd(Amp)]-fragment. Preference of the nucleosides was further evaluated by reacting 1 with two different nucleosides, and guanosine showed the highest selectivity among the series.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/molecules27072098/s1. Table S1: Hydrogen-bonding distances and angles; Figure S1: 1H NMR spectrum of 1 in CD3OD; Figure S2: 1H NMR spectrum of 1′ in CD3OD; Figure S3: 1H NMR spectrum of 2 in CD3CN; Figure S4: 1H NMR spectra of (a) reaction mixture of 1 and cytidine, (b) cytidine in CD3OD; Figure S5: 1H NMR spectra of (a) reaction mixture of 1 and 5-methyluridine, (b) 1, 5-methyluridine, and triethylamine, and (c) 5-methyluridine; cytidine in CD3OD; Figure S6: 1H NMR spectra of (a) reaction mixture of 1 and adenosine, (b) adenosine in CD3OD; Figure S7: 1H NMR spectra of (a) reaction mixture of 1 and guanosine, (b) guanosine in CD3OD.

Author Contributions

Conceptualization, R.M. and T.S.; methodology, R.M.; validation, R.M.; investigation, R.M. and Y.I.; resources, R.M., T.S. and Y.H.; writing—original draft, R.M.; writing—review and editing, R.M., T.S. and Y.H.; visualization, R.M.; project administration, R.M. All authors have read and agreed to the published version of the manuscript.

Funding

This work was partly funded by Sakigake Project 2020 of Kanazawa University. This work was partly supported by a Grant-in-aid for Scientific Research No. 19K15525 from MEXT, Japan.

Data Availability Statement

The crystallographic data are available from the Cambridge Crystallographic Data Centre (CCDC). Other data not presented in Supplementary Materials are available on request from the corresponding author.

Acknowledgments

Single-crystal X-ray analysis was conducted at the Institute of Molecular Science, supported by Nanotechnology Platform (Molecule and Material Synthesis) of the MEXT, Japan.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not applicable.

References

  1. Desiraju, G.R.; Steiner, T. The Weak Hydrogen Bond in Structural Chemistry and Biology; Oxford University Press: Oxford, UK, 2001. [Google Scholar] [CrossRef]
  2. Gilli, G.; Gilli, P. The Nature of Hydrogen Bond: Outline of a Comprehensive Hydrogen Bond Theory; Oxford University Press: Oxford, UK, 2009. [Google Scholar] [CrossRef] [Green Version]
  3. Burrows, A.D. Crystal Engineering Using Multiple Hydrogen Bonds. In Supramolecular Assembly via Hydrogen Bond I Structure and Bondings; Mingos, D.M.P., Ed.; Springer: Amsterdam, The Netherlands, 2004; Volume 108, pp. 55–96. [Google Scholar] [CrossRef]
  4. Braga, D.; Maini, L.; Polito, M.; Grepioni, F. Crystal Engineering Using Multiple Hydrogen Bonds. In Supramolecular Assembly via Hydrogen Bond II Hydrogen Bonding Interactions Between Ions: A Powerful Took in Molecular Crystal Engineering; Mingos, D.M.P., Ed.; Springer: Amsterdam, The Netherlands, 2004; Volume 111, pp. 1–32. [Google Scholar] [CrossRef]
  5. Mitsuhashi, R.; Suzuki, T.; Hosoya, S.; Mikuriya, M. Hydrogen-Bonded Supramolecular Structures of Cobalt(III) Complexes with Unsymmetrical Bidentate Ligands: Mer/fac Interconversion Induced by Hydrogen-Bonding Interactions. Cryst. Growth Des. 2017, 17, 207–213. [Google Scholar] [CrossRef]
  6. Mitsuhashi, R.; Pedersen, K.S.; Ueda, T.; Suzuki, T.; Bendix, J.; Mikuriya, M. Field-induced single-molecule magnet behavior in ideal trigonal antiprismatic cobalt(II) complexes: Precise geometrical control by a hydrogen-bonded rigid metalloligand. Chem. Commun. 2018, 54, 8869–8872. [Google Scholar] [CrossRef]
  7. Mitsuhashi, R.; Hosoya, S.; Suzuki, T.; Sunatsuki, Y.; Sakiyama, H.; Mikuriya, M. Hydrogen-bonding interactions and magnetic relaxation dynamics in tetracoordinated cobalt(II) single-ion magnets. Dalton Trans. 2019, 48, 395–399. [Google Scholar] [CrossRef] [PubMed]
  8. Mitsuhashi, R.; Hosoya, S.; Suzuki, T.; Sunatsuki, Y.; Sakiyama, H.; Mikuriya, M. Zero-field slow relaxation of magnetization in cobalt(II) single-ion magnets: Suppression of quantum tunneling of magnetization by tailoring the intermolecular magnetic coupling. RSC Adv. 2020, 10, 43472–43479. [Google Scholar] [CrossRef]
  9. Watson, J.D.; Crick, F.H.C. Molecular Structure of Nucleic Acids: A Structure for Deoxyribose Nucleic Acid. Nature 1953, 171, 737–738. [Google Scholar] [CrossRef] [PubMed]
  10. Nikolova, E.N.; Kim, E.; Wise, A.A.; O’Brien, P.J.; Andricioaei, I.; Al-Hashimi, H.M. Transient Hoogsteen base pairs in canonical duplex DNA. Nature 2011, 470, 498–502. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Rosenberg, B.; Camp, L.V.; Krigas, T. Inhibition of Cell Division in Escherichia coli by Electrolysis Products from a Platinum Electrode. Nature 1969, 205, 698–699. [Google Scholar] [CrossRef]
  12. Rosenberg, B.; Camp, L.V.; Trosko, J.E.; Mansour, V.H. Platinum Compounds: A New Class of Potent Antitumour Agents. Nature 1969, 222, 385–386. [Google Scholar] [CrossRef] [PubMed]
  13. Lovejoy, K.S.; Todd, R.C.; Zhang, S.; McCormick, M.S.; D’Aquino, J.A.; Reardon, J.T.; Sancar, A.; Giacomini, K.M.; Lippard, S.J. cis-Diammine(pyridine)chloroplatinum(II), a monofunctional platinum(II) antitumor agent: Uptake, structure, function, and prospects. Proc. Natl. Acad. Sci. USA 2008, 105, 8902–8907. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Todd, R.C.; Lippard, S.J. Structure of duplex DNA containing the cisplatin 1,2-{Pt(NH3)2}2+-d(GpG) cross-link at 1.77Å resolution. J. Inorg. Biochem. 2010, 104, 902–908. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Mitsuhashi, R.; Suzuki, T.; Sunatsuki, Y.; Kojima, M. Hydrogen-bonding interactions, geometrical selectivity and spectroscopic properties of cobalt(III) complexes with unsymmetrical tridentate amine-amidato-phenolato type ligands. Inorg. Chim. Acta 2013, 399, 131–137. [Google Scholar] [CrossRef]
  16. Mitsuhashi, R.; Ogawa, R.; Ishikawa, R.; Suzuki, T.; Sunatsuki, Y.; Kawata, S. Preparation, structures and properties of manganese complexes containing amine–(amido or amidato)–phenolato type ligands. Inorg. Chim. Acta 2016, 447, 113–120. [Google Scholar] [CrossRef]
  17. Schlosser, M. (Ed.) Organometallics in Synthesis: A Manual, 2nd ed.; Wiley: Hoboken, NJ, USA, 2004; 1126p. [Google Scholar]
  18. Hofmann, A.; Dahlenburg, L.; van Eldik, R. Cyclometalated analogues of platinum terpyridine complexes: Kinetic study of the strong sigma-donor cis and trans effects of carbon in the presence of a pi-acceptor ligand backbone. Inorg. Chem. 2003, 42, 6528–6538. [Google Scholar] [CrossRef] [PubMed]
  19. Odoko, M.; Okabe, N. (Adeninato-N9)[N-(2-aminoethyl)salicylideneiminato]palladium(II) 3.5-hydrate. Acta Crystallogr. Sect. E Struct. Rep. Online 2005, 61, m2670–m2673. [Google Scholar] [CrossRef] [Green Version]
  20. Sherman, S.E.; Lippard, S.J. Structural aspects of platinum anticancer drug interactions with DNA. Chem. Rev. 1987, 87, 1153–1181. [Google Scholar] [CrossRef]
  21. Gottlieb, H.E.; Kotlyar, V.; Nudelman, A. NMR chemical shifts of common laboratory solvents as trace impurities. J. Org. Chem. 1997, 62, 7512–7515. [Google Scholar] [CrossRef] [PubMed]
  22. Sheldrick, G.M. SHELXT—Integrated space-group and crystal-structure determination. Acta Crystallogr. Sect. A Found. Adv. 2015, 71, 3–8. [Google Scholar] [CrossRef] [Green Version]
  23. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. Sect. C Struct. Chem. 2015, 71, 3–8. [Google Scholar] [CrossRef]
Figure 1. Chemical structures of ligand precursors H2Amp and H2Apr.
Figure 1. Chemical structures of ligand precursors H2Amp and H2Apr.
Molecules 27 02098 g001
Figure 2. 1H NMR spectra of (a) 1, (b) 1′ in CD3OD, and (c) 2 in CD3CN.
Figure 2. 1H NMR spectra of (a) 1, (b) 1′ in CD3OD, and (c) 2 in CD3CN.
Molecules 27 02098 g002
Figure 3. Perspective views of (a) [Pd(Amp)Cl] in 1, and (b) [Pd(Amp)(CH3CN)] in 1′ (50% probability levels).
Figure 3. Perspective views of (a) [Pd(Amp)Cl] in 1, and (b) [Pd(Amp)(CH3CN)] in 1′ (50% probability levels).
Molecules 27 02098 g003
Figure 4. The asymmetric unit of 2 (50% probability levels). An intermolecular hydrogen bond is shown as magenta dashed line.
Figure 4. The asymmetric unit of 2 (50% probability levels). An intermolecular hydrogen bond is shown as magenta dashed line.
Molecules 27 02098 g004
Figure 5. Chemical structures of the nucleosides employed in this study. Hydrogen-bond donors and acceptors are shown in blue and red color.
Figure 5. Chemical structures of the nucleosides employed in this study. Hydrogen-bond donors and acceptors are shown in blue and red color.
Molecules 27 02098 g005
Figure 6. 1H NMR spectra of (a) the reaction mixture of 1 and cytidine, (b) cytidine in CD3OD. The resonances from a minor product and unreacted cytidine are marked with triangles and asterisks, respectively.
Figure 6. 1H NMR spectra of (a) the reaction mixture of 1 and cytidine, (b) cytidine in CD3OD. The resonances from a minor product and unreacted cytidine are marked with triangles and asterisks, respectively.
Molecules 27 02098 g006
Figure 7. Chemical structure of Pd-C (R indicates the ribose moiety). Hydrogen-bond donor and acceptor are shown in blue and red, respectively. Dashed lines in magenta indicate intramolecular hydrogen bonds.
Figure 7. Chemical structure of Pd-C (R indicates the ribose moiety). Hydrogen-bond donor and acceptor are shown in blue and red, respectively. Dashed lines in magenta indicate intramolecular hydrogen bonds.
Molecules 27 02098 g007
Figure 8. 1H NMR spectra of (a) the reaction mixture of 1 and 5-methyluridine, (b) 1, 5-methyluridine, and triethylamine, and (c) 5-methyluridine in CD3OD. The resonances for unreacted 5-methyluridine are marked with asterisks.
Figure 8. 1H NMR spectra of (a) the reaction mixture of 1 and 5-methyluridine, (b) 1, 5-methyluridine, and triethylamine, and (c) 5-methyluridine in CD3OD. The resonances for unreacted 5-methyluridine are marked with asterisks.
Molecules 27 02098 g008
Figure 9. Three possible coordination modes for Pd-A (R indicates the ribose moiety). (a) N1-coordination mode, (b) N3-coordination mode, and (c) N7-coordination mode. Hydrogen bonds are shown in magenta.
Figure 9. Three possible coordination modes for Pd-A (R indicates the ribose moiety). (a) N1-coordination mode, (b) N3-coordination mode, and (c) N7-coordination mode. Hydrogen bonds are shown in magenta.
Molecules 27 02098 g009
Figure 10. 1H NMR spectra of (a) the reaction mixture of 1 and adenosine, and (b) adenosine in CD3OD. The resonances of aryl-H atoms for two minor products and adenosine are marked in the spectrum (Pd-A2: open triangle; Pd-A3: open square; adenosine: asterisk).
Figure 10. 1H NMR spectra of (a) the reaction mixture of 1 and adenosine, and (b) adenosine in CD3OD. The resonances of aryl-H atoms for two minor products and adenosine are marked in the spectrum (Pd-A2: open triangle; Pd-A3: open square; adenosine: asterisk).
Molecules 27 02098 g010
Figure 11. Chemical structure of Pd-G anion (R indicates the ribose moiety). Hydrogen-bond donor and acceptor are shown in blue and red, respectively. Dashed lines in magenta indicate intramolecular hydrogen bonds.
Figure 11. Chemical structure of Pd-G anion (R indicates the ribose moiety). Hydrogen-bond donor and acceptor are shown in blue and red, respectively. Dashed lines in magenta indicate intramolecular hydrogen bonds.
Molecules 27 02098 g011
Figure 12. 1H NMR spectra of (a) the reaction mixture of 1 and guanosine, and (b) guanosine in CD3OD.
Figure 12. 1H NMR spectra of (a) the reaction mixture of 1 and guanosine, and (b) guanosine in CD3OD.
Molecules 27 02098 g012
Figure 13. 1H NMR spectra of 1 with two different nucleosides in CD3OD. (a) Adenosine and cytidine, (b) adenosine and guanosine, and (c) guanosine and cytidine. The resonances attributed to the major and minor products for each nucleoside are indicated with solid circles and triangles, respectively (the resonances for Amp2− ligand for each species are not marked). Pd-A, Pd-G, and Pd-C species are shown in blue, green, and red, respectively. Asterisks indicate unreacted nucleoside.
Figure 13. 1H NMR spectra of 1 with two different nucleosides in CD3OD. (a) Adenosine and cytidine, (b) adenosine and guanosine, and (c) guanosine and cytidine. The resonances attributed to the major and minor products for each nucleoside are indicated with solid circles and triangles, respectively (the resonances for Amp2− ligand for each species are not marked). Pd-A, Pd-G, and Pd-C species are shown in blue, green, and red, respectively. Asterisks indicate unreacted nucleoside.
Molecules 27 02098 g013
Table 1. Selected bond parameters for 1.
Table 1. Selected bond parameters for 1.
Atom–AtomLength/ÅAtom–Atom–AtomAngle/°
Pd(1)–O(1)1.9877 (13)N(1)–Pd(1)–O(1)93.70 (6)
Pd(1)–N(1)1.9647 (14)N(1)–Pd(1)–N(2)83.87 (6)
Pd(1)–N(2)2.0264 (16)O(1)–Pd(1)–N(2)177.15 (5)
Pd(1)–Cl(1)2.3389 (4)N(1)–Pd(1)–Cl(1)176.50 (4)
O(1)–Pd(1)–Cl(1)89.79 (4)
N(2)–Pd(1)–Cl(1)92.64 (4)
Table 2. Selected bond parameters for 1′.
Table 2. Selected bond parameters for 1′.
Atom–AtomLength/ÅAtom–Atom–AtomAngle/°
Pd(1)–O(1)1.9771 (19)N(1)–Pd(1)–O(1)94.73 (8)
Pd(1)–N(1)1.956 (2)N(1)–Pd(1)–N(2)82.77 (9)
Pd(1)–N(2)2.025 (2)O(1)–Pd(1)–N(2)176.79 (8)
Pd(1)–N(3)2.027 (2)N(1)–Pd(1)–N(3)176.99 (8)
O(1)–Pd(1)–N(3)88.02 (8)
N(2)–Pd(1)–N(3)94.44 (9)
Table 3. Selected bond parameters for 2.
Table 3. Selected bond parameters for 2.
Atom–AtomLength/ÅAtom–Atom–AtomAngle/°
Pd(1)–C(5)1.982 (5)N(1)–Pd(1)–C(5)81.8 (2)
Pd(1)–N(1)1.974 (5)N(1)–Pd(1)–N(2)92.2 (2)
Pd(1)–N(2)2.144 (4)C(5)–Pd(1)–N(2)173.2 (3)
Pd(1)–N(3)1.986 (5)N(1)–Pd(1)–N(3)175.65 (17)
Pd(2)–C(17)1.981 (6)C(5)–Pd(1)–N(3)95.1 (2)
Pd(1)–N(4)1.998 (4)N(2)–Pd(1)–N(3)90.72 (19)
Pd(1)–N(5)2.121 (5)N(4)–Pd(2)–C(17)81.8 (2)
Pd(1)–N(6)2.013 (5)N(4)–Pd(2)–N(5)94.33 (18)
C(17)–Pd(2)–N(5)173.84 (19)
N(4)–Pd(2)–N(6)176.3 (2)
C(17)–Pd(2)–N(6)95.3 (2)
N(5)–Pd(2)–N(6)88.41 (19)
Table 4. Crystallographic data and refinement parameters of 1′ and 2.
Table 4. Crystallographic data and refinement parameters of 1′ and 2.
Complex11′2
Empirical formulaC15H28ClN3O3PdC15H20N4O2PdC12H15N3O2Pd
Formula weight440.25394.75339.67
Crystal systemMonoclinicMonoclinicTriclinic
Crystal dimensions/mm0.19 × 0.12 × 0.060.21 × 0.15 × 0.090.11 × 0.07 × 0.04
Space groupP21/cP21/cP−1
a11.8363 (3)13.3692 (5)8.7885 (4)
b8.7436 (2)10.9488 (4)10.8746 (5)
c18.3451 (5)11.1827 (4)13.2005 (5)
α/° 96.688 (4)
β/°104.732 (3)94.084 (3)90.840 (3)
γ/° 104.073 (4)
V31836.15 (8)1632.73 (10)1214.19 (9)
Z444
T/K100 (2)100 (2)100 (2)
ρcalcd/g·cm−31.5931.6061.858
µ/mm−11.1731.1491.526
F(000)904800680
2θmax/◦555555
No. of reflections measured17,46911,61617,415
No. of independent reflections4207 (Rint = 0.0244)3737 (Rint = 0.0390)5543 (Rint = 0.1488)
Data/restraints/parameters4207/2/2223737/0/2035543/18/329
R1 1 [I > 2.00 σ(I)]0.02170.03200.0555
wR2 2 (all reflections)0.05580.08030.1348
Goodness of fit indicator1.0761.0880.906
Highest peak, deepest hole/e Å−30.445, −0.4390.627, −1.1561.646, −1.379
CCDC deposition number215300521530062153007
1R1 = Σ||Fo| − |Fc||/Σ|Fo|, 2 wR2 = [Σ(w(Fo2 − Fc2)2)/Σw(Fo2)2]1/2.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mitsuhashi, R.; Imai, Y.; Suzuki, T.; Hayashi, Y. Selective Formation of Intramolecular Hydrogen-Bonding Palladium(II) Complexes with Nucleosides Using Unsymmetrical Tridentate Ligands. Molecules 2022, 27, 2098. https://doi.org/10.3390/molecules27072098

AMA Style

Mitsuhashi R, Imai Y, Suzuki T, Hayashi Y. Selective Formation of Intramolecular Hydrogen-Bonding Palladium(II) Complexes with Nucleosides Using Unsymmetrical Tridentate Ligands. Molecules. 2022; 27(7):2098. https://doi.org/10.3390/molecules27072098

Chicago/Turabian Style

Mitsuhashi, Ryoji, Yuya Imai, Takayoshi Suzuki, and Yoshihito Hayashi. 2022. "Selective Formation of Intramolecular Hydrogen-Bonding Palladium(II) Complexes with Nucleosides Using Unsymmetrical Tridentate Ligands" Molecules 27, no. 7: 2098. https://doi.org/10.3390/molecules27072098

Article Metrics

Back to TopTop