Next Article in Journal
Diet-Induced Metabolic Dysfunction of Hypothalamic Nutrient Sensing in Rodents
Next Article in Special Issue
Supramolecular Hydrogen Bonding Assembly from Non-Coplanar Aromatic Tetra-1H-Pyrazoles with Crystallization-Induced Emission (CIE)
Previous Article in Journal
The Preferential Use of Anakinra in Various Settings of FMF: A Review Applied to an Updated Treatment-Related Perspective of the Disease
Previous Article in Special Issue
Type I–IV Halogen⋯Halogen Interactions: A Comparative Theoretical Study in Halobenzene⋯Halobenzene Homodimers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhancement of the Water Affinity of Histidine by Zinc and Copper Ions

1
School of Physics, East China University of Science and Technology, Shanghai 200237, China
2
Shanghai Applied Radiation Institute, Shanghai University, Shanghai 200444, China
3
Zhangjiang Laboratory, Shanghai Advanced Research Institute, Chinese Academy of Sciences, Shanghai 201210, China
4
Shanghai Institute of Applied Physics, Chinese Academy of Sciences, Shanghai 201800, China
5
Wenzhou Institute, University of Chinese Academy of Sciences, Wenzhou 325000, China
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(7), 3957; https://doi.org/10.3390/ijms23073957
Submission received: 16 February 2022 / Revised: 27 March 2022 / Accepted: 30 March 2022 / Published: 2 April 2022
(This article belongs to the Special Issue Non-covalent Interaction)

Abstract

:
Histidine (His) is widely involved in the structure and function of biomolecules. Transition-metal ions, such as Zn2+ and Cu2+, widely exist in biological environments, and they are crucial to many life-sustaining physiological processes. Herein, by employing density function calculations, we theoretically show that the water affinity of His can be enhanced by the strong cation–π interaction between His and Zn2+ and Cu2+. Further, the solubility of His is experimentally demonstrated to be greatly enhanced in ZnCl2 and CuCl2 solutions. The existence of cation–π interaction is demonstrated by fluorescence, ultraviolet (UV) spectroscopy and nuclear magnetic resonance (NMR) experiments. These findings are of great importance for the bioavailability of aromatic drugs and provide new insight for understanding the physiological functions of transition metal ions.

1. Introduction

Histidine (His) is one of the most essential and naturally occurring aromatic amino acids. As one of the building blocks of proteins, it is commonly involved in the structure and function of proteins, since its side-chain contains an imidazole ring which can easily participate in π–π [1,2], hydrogen bond [3,4], coordinate bond [5,6,7,8] and cation–π interactions [9,10,11]. For example, the coordination of ligand to the imidazole ring of His can inhibit the enzymic activity of tyrosine hydroxylase [12] and π–π stacking between His and Tyr is associated with the mitogenicity function of lectin [13]. Besides, histidine-rich peptides can be incorporated into polymers, liposomes, and proteins, including virus-like particles [14,15,16].
In many proteins, His is coordinated with transition metal ions, among which zinc and copper are the two with the most occurrence. This coordination is critical to the structures and functions of biomolecules [17,18,19,20,21]. For instance, the carboxylate–histidine–zinc triad is frequently found in zinc proteins and is important in the function of these metalloproteins [22]. The riboflavin binding protein (RBP) can bind copper and form a distinct well-ordered type II site under dialysis conditions [19]. In addition, the zinc–histidine complex had been used as a zinc supplement and was shown better absorption than zinc sulfate in humans [23,24]. The copper–histidine complex has been found in human blood, and has been applied to treat Menkes disease [25].
Most biochemical processes occur in aqueous solution. The dispersion behavior of biomolecules in water is important for their participation in physical, chemical, and biological processes [26,27,28,29,30]. Previous studies on the coordination of His with transition metal ions are mainly focused on the specific coordination modes and related biological functions. Whereas, studies on some general features of this coordination, such as the hydrophilicity of His, are fewer. By the binding of zinc or copper, the hydrophilicity of His will be modified, which will have significance for a wide range of problems, such as protein folding [31,32], structure stabilization [33,34]. The effect on the hydrophilicity of His by the binding of zinc/copper can be studied through the solubility change for His and complex Zn2+/Cu2+-His. Shi et al. studied the solubility of tryptophan (Trp) in transition metal ion salt solutions and found that the solubility of Trp is dramatically increased [35].
In this paper, by employing DFT calculations, we investigate the interaction between His and transition metal ions Zn2+ and Cu2+. Experiments on the solubility of His in ZnCl2 and CuCl2 solutions were then performed to verify the results of theoretical calculations. First principles calculations and solubility experiments strongly support that the strong cation–π interaction between His and transition metal ions Zn2+ and Cu2+ greatly affects the water affinity of His. Finally, the cation–π interaction between His and Zn2+/Cu2+ was demonstrated by fluorescence, ultraviolet (UV) spectroscopy and nuclear magnetic resonance (NMR) experiments.

2. Results and Discussion

To investigate the interaction between His and transition metal ions Zn2+ and Cu2+, we first calculated the interaction energies between the imidazole ring in His with and without Zn2+ binding and the nearest neighbor water molecule by density functional theory (DFT). As shown in Figure 1b, when Zn2+ interacts with His, the interaction energy between the imidazole ring in His and the nearest water molecule is about −12.5 kcal·mol−1, which is much larger than the corresponding interaction energy without Zn2+ binding (−3.5 kcal·mol−1, Figure 1a). In addition, we also calculated the interaction energy between the imidazole ring in His and the nearest water molecule with Cu2+ binding, which is about −13.0 kcal·mol−1 (Figure 1c). These results show that both Zn2+ and Cu2+ greatly increase the water affinity of His to about the same degree.
It should be pointed out that in the most stable structures of Zn2+–His and Cu2+–His, Zn2+ and Cu2+ interact with the side of the imidazole ring together with the amino N atom and carbonyl O atom rather than at the top of the imidazole ring. The interaction energy of this ring/N/O tridentate coordination is 20.3 kcal·mol−1 larger than that of the structure in which Zn2+ stays at the top of the imidazole ring (Figure S1). Rimola et al. also pointed out that the configuration of Cu2+−His with Cu2+ at the top of the imidazole ring was not obtainable by first principle calculations [36]. Besides, Sastry et al. calculated the interaction between multiple cations and imidazole, finding that cations show affinity only toward the heteroatom N atom instead of the whole π–face of the imidazole ring [9]. Thus, the configuration of Zn2+/Cu2+ stays at the top of the imidazole ring, which interacts with a traditional cation–π interaction way, could not be the real complex in the experiments. For the configuration of ring/N/O tridentate coordination, it is expected that there is also a cation–π interaction between Zn2+/Cu2+ and His, because the π electron of imidazole ring and valence electron of Zn2+/Cu2+ are both involved in the HOMO-1 orbital (Figure S2) and the electron distribution of the whole imidazole ring is affected by the Zn2+ binding (Figure S3). To further analyze the cation−π interactions, we provide EDA calculations on the Zn2+/Cu2+−His complex. EDA can divide the total interaction energy into several interaction terms with physical meaning and is widely used for analyzing various intermolecular interactions [37,38]. We divided the optimized structure of Zn2+/Cu2+−His into two parts, i.e., the Zn2+/Cu2+ ion and His, and then performed single-point calculations at the B3LYP/TZP level of theory in the framework of DFT by using the ADF program [39]. The total interaction energies between Zn2+/Cu2+ and His were decomposed into Pauli repulsion, orbital interaction and electrostatic interaction (Table S1). The EDA results show that the electrostatic interactions and orbital interactions approximately equally contribute to the interaction energy, and the Pauli repulsion affords the main repulsive energy.
We also would like to note that the geometric structures of the Zn2+–His–water and Cu2+–His–water systems were optimized based on the water molecule forming a hydrogen bond with the imidazole ring at the C–H site (Figure 1b,c). In fact, water can interact with the imidazole ring of Zn2+–His and Cu2+–His complexes at other sites, for example, N−H of the imidazole ring (shown in Figure S4), which also indicate that the water affinity of imidazole ring is increased. Since N−H is not as hydrophobic as C−H in His, the increase of water affinity of N−H is not as critical as C−H for the water affinity enhancement of the whole complex.
The water affinity is well related to solubility. A weak interaction with water indicates that hydration of the compound is unfavorable, contributing to poor water solubility. Because the water affinities of Zn2+–His and Cu2+–His are much larger than that of His, we expected that the solubilities of His in ZnCl2 and CuCl2 solutions would both dramatically increase. We then performed experiments on the solubilities of His (SHis) in ZnCl2 and CuCl2 aqueous solutions. The solubility of His in ZnCl2 solution increases as much as that in CuCl2 solution (Figure 1d). In 0.4 M ZnCl2 and CuCl2 solutions, the solubility of His can increase to more than 5 times that in pure water. This is consistent with the first principles calculations of the interaction energy between the aromatic ring structure in His and the nearest water molecule. The results of His were then compared with previous work on Trp to show the enhanced water affinity of His by Zn2+ is nontrivial. Previous work on Trp [35] showed that the solubility of Trp in ZnCl2 solution slightly increases compared with the solubility in water (17.1 mg·mL−1 in 0.5 M ZnCl2 solution vs. 11.4 mg·mL−1 in pure water, Table 1). These results indicate that the aromatic amino acid His has a binding preference for Zn2+ compared with other amino acids and the interaction between them can greatly enhance the water affinity of His.
We would like to point out that there is no obvious difference when different orders of adding cation ions and His are employed. Both orders will lead to the same result that the solubility of His is enhanced significantly. This is a major difference between His in this work and Trp in previous work [35], where the order of adding cations and Trp show totally different results for the solubility enhancement of Trp.
The solubility behavior of His with respect to the concentrations of ZnCl2 and CuCl2 solutions was then investigated. As shown in Figure 1e and Figure S5, the solubility behavior can be well fitted by S A = A M C M + S A 0 , where S A and S A 0 are the solubilities of amino acid A in the salt solution and pure water, A M is the water affinity factor of amino acid A induced by metal ion M2+, and C M is the concentration of M2+. A M has a distinct physical meaning that for every M salt concentration increase, the solubility of the amino acid increases by A M M. For His, A Cu = 4.86 and A Zn = 2.80, which are both much greater than the water affinity factor of Trp induced by Cu2+ ( A Cu = 0.46) [35].
It should be noted that different Zn2+–His and Cu2+–His complexes exist at different pH values in experiments [40]. The structure that we used for the DFT calculations was referred to as the M(HL)2+ species in refs [40,41]. Under weak acid conditions, this structure of tridentate coordination is the most accepted one [42,43]. Indeed, the pH values of the ZnCl2 and CuCl2 solutions used in the experiments correspond to the weak acid condition (5.24–5.75 for the ZnCl2 solutions and 3.04–5.75 for the CuCl2 solutions) (Tables S2 and S3).
To verify that the increase of the solubility mainly comes from the cation–π interaction between the transition-metal ion and the imidazole ring in His, the solubility of the non-aromatic amino acid glycine (Gly) was measured under the same conditions. In Figure S5, it can be found that the solubilities of Trp and His both dramatically increased in CuCl2 solutions, while the solubility of Gly only showed a slight increase. This result strongly supports that the enhanced solubility of His can be mainly attributed to the imidazole ring in His. Zn2+ is expected to show the same behavior.
Fluorescence and ultraviolet (UV) absorption spectral experiments were then performed to provide evidence for the cation–π interaction between the metal ion and the imidazole ring in His. In Figure 2a, the fluorescence spectrum of His excited at 360 nm has an emission peak at 448 nm, which is assigned to the conjugated double bonds of the imidazole ring that can easily generate the π–π* transition [44]. Compared with the fluorescence intensity of His in water, the intensity of His in 50 mM CuCl2 solution markedly decreased, indicating that the conjugated double bonds of the imidazole ring in His were greatly affected in CuCl2 solution. The fluorescence spectrum of His in ZnCl2 solution is also shown in Figure 2a. Again, the fluorescence intensities of His and ZnCl2 were quenched when the Zn2+–His complex was formed, but to a smaller degree. The reason for the difference of the quenching degree in fluorescence is beyond the scope of this study. These results indicate that in ZnCl2 and CuCl2 solutions, Zn2+ and Cu2+ directly interact with the imidazole ring in His and decrease the fluorescence intensity of His. The imidazole ring of His exhibits absorption in the lower UV region (220 nm) [45]. Here, we observed that the UV absorption spectrum of His was also affected by the cation–π interaction between the imidazole ring in His and Cu2+ in solution (Figure S6), which is important evidence for the existence of the cation–π interaction. Overall, the fluorescence and UV absorption spectral experiments show the existence of the cation–π interaction between the imidazole ring in His and the metal ion, which is consistent with the theoretical prediction.
Nuclear magnetic resonance (NMR) experiments on C12 also showed that Cu2+–His had a clear chemical shift at 130 ppm, which is attributed to the imidazole ring in His (Figure 2b) [46]. Zn2+–His also showed a chemical shift at 130 ppm, although with a weaker signal (data not shown).
The structure of His did not change much after the binding of the transition metal ions Zn2+ and Cu2+. Infrared (IR) spectra of the dried samples of His in ZnCl2 and CuCl2 solutions were similar to that in pure water (Figure S7) [47,48].
The enhancement of SHis in ZnCl2 and CuCl2 solutions does not come from the pH effect induced by hydrolysis of Zn2+ or Cu2+. As shown in Tables S2 and S3, the pH values of the ZnCl2 and CuCl2 solutions both slightly decreased (from 6.0 to 5.2 for Zn2+ and from 4.5 to 3.1 for Cu2+), indicating that the pH effect does not greatly contribute to the increased solubility of His.

3. Materials and Methods

3.1. Computational Methods

The B3LYP functional [49] in the framework of DFT is used to calculate the Zn2+–His–H2O and Cu2+–His–H2O systems with Gaussian 09 package [50]. Geometric structures were first optimized by Berny algorithm [51] with the convergence criteria of a maximum step size of 0.0018 au and a root mean square force of 0.0003 a.u. A hybrid pseudo potential LanL2DZ is employed to calculate Cu2+ and Zn2+, while other atoms are calculated at the 6–31+G(d,p) basis set level (see detailed methods in the Supplementary Material (SM)).
Interaction energies between water molecule and aromatic ring of amino acid (AA) with and without metal ion (M) binding are represented as ΔEi, which can be calculated as,
Δ E i = E Total E complex E w ,
where, E Total , E complex , and E w are the single-point energies of His–M2+–H2O (or His–H2O), complex His–M2+ (or His), and H2O, respectively.
All the calculations are performed in vacuum condition. There is no other water molecule and ion except the nearest water and the bound ion we considered explicitly.

3.2. Experiments Materials

L–Histidine (His, 99%) and Glycine (Gly, 99%) were purchased from Sigma-Aldrich, Shanghai, China. Zinc(II) chloride (ZnCl2, 98%) was purchased from J&K Scientific Ltd, Beijing, China. Copper(II) chloride dihydrate (CuCl2·2H2O) (99%) was purchased from Sinopharm Chemical Reagent Co., Ltd, Beijing, China. All samples were used without purification and preprocessing. All salt solutions and other aqueous solutions were prepared with 18.2 MΩ, 3 ppb TOC Milli–Q water (Millipore, Burlington, MA, USA).

3.3. Solubility Measurement

Histidine (or Glycine) powder was added into pure water (or into the given 0.01–0.4 M CuCl2 and ZnCl2 solutions) with constant shaking until apparent saturation (with some insoluble His (or Gly) powder appeared), and then the solution was kept continuously stirring for 24 h in a thermostat at 25 °C. Afterward, the solution was centrifuged at 25 °C to remove insoluble His (or Gly) powder. Then, approximately 1 mL of solution was withdrawn by a pipette from the supernatant phase and transferred to a clean and weighed centrifuge tube. These centrifuge tubes were then transferred to liquid nitrogen for freezing and lyophilized overnight in a vacuum flask at 0.125 mbar and −58 °C in a freeze-dryer (Virtis Freezer Dryer, New York, NY, USA). The drying process was repeated until a constant mass reading was achieved. The solubility of His (or Gly) was calculated by the mass value difference of the centrifuge tube after removing salt mass. The data reported in this work were ensured by measurement of solubility for at least three replicate experiments at all compositions (Tables S2 and S3). Based on our measurement strategy, the solubility of His in pure water is 41.9 mg·mL−1, consistent with previous reports [52].

3.4. Measurement of pH

The pH values were measured by SevenCompact™ S220 pH meter (METTLER TOLEDO, Zurich, Switzerland) (pH = 0~14).

3.5. UV Spectroscopy

UV absorption spectra of His, CuCl2, and His (CuCl2) solutions were recorded on a U–3100 spectrophotometer (Hitachi, Tokyo, Japan). The concentration for His used in this experiment is 30 mg·mL−1, for CuCl2 is 50 mM.

3.6. Fluorescence Spectrofluorophotometer

Excitation and photoluminescence (PL) spectra were measured with a Hitachi 7000 fluorescence spectrophotometer (Tokyo, Japan) and emission slit width of 10 mm was used to record fluorescence spectra, and the fluorescence spectra of the work were recorded with λ e x / λ e m = 360 nm/440 nm. The thickness of all liquid sample cells is 10 mm.

3.7. IR Spectra

IR spectra from 4000 cm−1 to 500 cm−1 of His, Zn2+–His, and Cu2+–His powder were measured using a Bio-Rad FTIR spectrometer FTS165 (Benton, ME, USA) equipped with resolution of 4 cm−1. The drying His, Zn2+–His, and Cu2+–His powder were obtained by freezing and lyophilized from corresponding saturated solution.

3.8. Solid-State NMR Spectroscopy

Solid-state NMR experiments were carried out on a wide-bore Bruker AVANCE-600 spectrometer (14.1 T) and a Bruker DSX-400 spectrometer (Karlsruhe, Germany) on 4-mm triple-resonance MAS probes. The drying His, Zn2+–His, and Cu2+–His powders were obtained by freezing and lyophilized from corresponding saturated solutions.

4. Conclusions

In summary, by first principles calculations, we have shown that the cation–π interaction between Zn2+ and His is very strong, which can enhance the water affinity of His to a comparable extent to Cu2+. Theoretical studies showed that the strong cation–π interaction between Zn2+/Cu2+ and His is the key. This cation–π interaction modifies the electronic distribution of the imidazole ring in His and significantly enhances the water affinity of His. Further EDA analysis showed that this cation–π interaction is approximately contributed equally by electrostatic interactions and orbital interactions. We also performed solubility experiments, which showed that the solubilities of His in ZnCl2 and CuCl2 solutions can reach more than 5 times that of His in pure water.
The results highlight that the solubility enhancement of many imidazole derivatives by Zn2+ and Cu2+ may be a general phenomenon and needs to attract more attention in different research fields, such as drug chemistry and colloid chemistry. These findings will enrich the understanding of the interactions between metal ions and biomolecules, and provide new insight into the physiological functions of multivalent metal ions.
We also would like to note that, our previous study on the solubility of Trp showed that Zn2+ is relatively weak in affecting the hydrophobicity of Trp by cation–π interaction, which is about one third of the solubility enhancement of Cu2+ [35]. Combined with this study, we found that the cation–π interaction is sensitive to the specific interaction environment, providing a possible scheme for the selectivity mechanism of biomolecules. Actually, Tu et al. recently showed that the cation–π interaction is the origin of the selectivity mechanism of calcium and magnesium in phosphotyrosine, demonstrating the sensitivity of cation–π interaction to different ion species with identical charges [53]. Considering the common binding mode of zinc-histidine in biology, it is expected histidine will participate in the selectivity mechanism of transition metal ions.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms23073957/s1. Reference [54] is cited in Supplementary Materials.

Author Contributions

Conceptualization, H.F. and G.S.; methodology, J.Z. and M.L.; Data curation and formal analysis, Y.S. and J.Z.; investigation, Y.S. and J.Z.; resources, H.F. and M.W.; writing—original draft preparation, Y.S.; writing—review and editing, H.F., G.S. and H.Z.; project administration, H.F. and M.W. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Natural Science Foundation of China (Grant Nos. 11974366 and 62075225), the Fundamental Research Funds for the Central Universities, China, the National Science Fund for Outstanding Young Scholars (Grant No. 11722548), the Key Research Program of Frontier Sciences of the Chinese Academy of Sciences (No. QYZDJ-SSW-SLH053), and the National Defense Science and Technology Innovation Special Zone Project.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We thank Xiaoling Lei, Shiqi Sheng, Shanshan Liang for their constructive suggestions and helpful discussions. We are also thankful for the discussions with the staff from the BL06B beamline of the Shanghai Synchrotron Radiation Facilities(SSRF). We thank the Shanghai Supercomputer Center of China. We also thank Tim Cooper for editing the English text of a draft of this manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sundberg, R.J.; Martin, R.B. Interactions of histidine and other imidazole derivatives with transition metal ions in chemical and biological systems. Chem. Rev. 1974, 74, 471–517. [Google Scholar] [CrossRef]
  2. Wang, L.; Sun, N.; Terzyan, S.; Zhang, X.; Benson, D.R. A histidine/tryptophan π-stacking interaction stabilizes the heme-independent folding core of microsomal apocytochrome b5 relative to that of mitochondrial apocytochrome b5. Biochemistry 2006, 45, 13750–13759. [Google Scholar] [CrossRef] [PubMed]
  3. Henry, M. Thermodynamics of hydrogen bond patterns in supramolecular assemblies of water molecules. ChemPhysChem 2002, 3, 607–616. [Google Scholar] [CrossRef]
  4. Graham, B.; Comba, P.; Hearn, M.T.W.; Spiccia, L. An examination of the binding behavior of histidine-containing peptides with immobilized metal complexes derived from the macrocyclic ligand, 1,4,7-triazacyclononane. J. Biol. Inorg. Chem. 2007, 12, 11–21. [Google Scholar] [CrossRef]
  5. Miessler, G.L.; Tarr, D.A. Inorganic Chemistry, 3rd ed.; Pearson Prentice Hall: Upper Saddle River, NJ, USA, 2003. [Google Scholar]
  6. Ginotra, Y.P.; Kulkarni, P.P. Solution structure of physiological Cu(His)2: Novel considerations into imidazole coordination. lnorg. Chem. 2009, 48, 7000–7002. [Google Scholar] [CrossRef]
  7. Annual Review of BiophysicsWatly, J.; Simonovsky, E.; Wieczorek, R.; Barbosa, N.; Miller, Y.; Kozlowski, H. Insight into the coordination and the binding sites of Cu2+ by the histidyl-6-tag using experimental and computational tools. lnorg. Chem. 2014, 53, 6675–6683. [Google Scholar]
  8. Milorey, B.; Malyshka, D.; Schweitzer-Stenner, R. pH dependence of ferricytochrome c conformational transitions during binding to cardiolipin membranes: Evidence for histidine as the distal ligand at neutral pH. J. Phys. Chem. Lett. 2017, 8, 1993–1998. [Google Scholar] [CrossRef]
  9. Reddy, A.S.; Sastry, G.N. Cation [M = H+, Li+, Na+, K+, Ca2+, Mg2+, NH4+, and NMe4+] interactions with the aromatic motifs of naturally occurring amino acids: A theoretical study. J. Phys. Chem. A 2005, 109, 8893–8903. [Google Scholar] [CrossRef]
  10. Dougherty, D.A. Cation-π interactions in chemistry and biology: A new view of benzene, Phe, Tyr, and Trp. Science 1996, 271, 163–168. [Google Scholar] [CrossRef]
  11. Liao, S.; Du, Q.; Meng, J.; Pang, Z.; Huang, R. The multiple roles of histidine in protein interactions. Chem. Cent. J. 2013, 7, 44. [Google Scholar] [CrossRef] [Green Version]
  12. Martínez, A. Evidence for a functionally important histidine residue in human tyrosine hydroxylase. Amino Acids 1995, 9, 285–292. [Google Scholar] [CrossRef]
  13. Covés-Datson, E.M.; King, S.R.; Legendre, M.; Swanson, M.D.; Gupta, A.; Claes, S.; Meagher, J.L.; Boonen, A.; Zhang, L.; Kalveram, B.; et al. Targeted disruption of π–πstacking in Malaysian banana lectin reduces mitogenicity while preserving antiviral activity. Sci. Rep. 2021, 11, 656. [Google Scholar] [CrossRef]
  14. Chang, K.-L.; Higuchi, Y.; Kawakami, S.; Yamashita, F.; Hashida, M. Development of lysine–histidine dendron modified chitosan for improving transfection efficiency in HEK293 cells. J. Control. Release 2011, 156, 195–202. [Google Scholar] [CrossRef]
  15. Gu, J.; Wang, X.; Jiang, X.; Chen, Y.; Chen, L.; Fang, X.-l.; Sha, X. Self-assembled carboxymethyl poly (L-histidine) coated poly +²-amino ester)/DNA complexes for gene transfection. Biomaterials 2012, 33, 644–658. [Google Scholar] [CrossRef]
  16. Ferrer-Miralles, N.; Corchero, J.L.; Kumar, P.; Cedano, J.A.; Gupta, K.C.; Villaverde, A.; Vazquez, E. Biological activities of histidine-rich peptides; merging biotechnology and nanomedicine. Microb. Cell Fact. 2011, 10, 101. [Google Scholar] [CrossRef] [Green Version]
  17. Berg, J.M.; Shi, Y. The galvanization of biology: A growing appreciation for the roles of zinc. Science 1996, 271, 1081–1085. [Google Scholar] [CrossRef]
  18. Hoch, E.; Lin, W.; Chai, J.; Hershfinkel, M.; Fu, D.; Sekler, I. Histidine pairing at the metal transport site of mammalian ZnT transporters controls Zn2+ over Cd2+ selectivity. Proc. Natl. Acad. Sci. USA 2012, 109, 7202. [Google Scholar] [CrossRef] [Green Version]
  19. Smith, S.R.; Bencze, K.Z.; Russ, K.A.; Wasiukanis, K.; Benore-Parsons, M.; Stemmler, T.L. Investigation of the Copper Binding Site and the Role of Histidine as a Ligand in Riboflavin Binding Protein. lnorg. Chem. 2008, 47, 6867–6872. [Google Scholar] [CrossRef] [Green Version]
  20. Higuchi, Y.; Suzuki, T.; Arimori, T.; Ikemura, N.; Mihara, E.; Kirita, Y.; Ohgitani, E.; Mazda, O.; Motooka, D.; Nakamura, S.; et al. Engineered ACE2 receptor therapy overcomes mutational escape of SARS-CoV-2. Nat. Commun. 2021, 12, 3802. [Google Scholar] [CrossRef]
  21. Thompsett, A.R.; Abdelraheim, S.R.; Daniels, M.; Brown, D.R. High Affinity Binding between Copper and Full-length Prion Protein Identified by Two Different Techniques*. J. Biol. Chem. 2005, 280, 42750–42758. [Google Scholar] [CrossRef] [Green Version]
  22. Christianson, D.W.; Alexander, R.S. Carboxylate-histidine-zinc interactions in protein structure and function. J. Am. Chem. Soc. 1989, 111, 6412–6419. [Google Scholar] [CrossRef]
  23. Schölmerich, J.; Krauss, E.; Wietholtz, H.; Köttgen, E.; Löhle, E.; Gerok, W. Bioavailability of zinc from zinc-histidine complexes. II Studies on patients with liver cirrhosis and the influence of the time of application. Am. J. Clin. Nutr. 1987, 45, 1487–1491. [Google Scholar] [CrossRef]
  24. Schölmerich, J.; Freudemann, A.; Köttgen, E.; Wietholtz, H.; Steiert, B.; Löhle, E.; Häussinger, D.; Gerok, W. Bioavailability of zinc from zinc-histidine complexes. I. Comparison with zinc sulfate in healthy men. Am. J. Clin. Nutr. 1987, 45, 1480–1486. [Google Scholar] [CrossRef]
  25. Sarkar, B. Treatment of Wilson and Menkes Diseases. Chem. Rev. 1999, 99, 2535–2544. [Google Scholar] [CrossRef]
  26. Bahng, J.H.; Yeom, B.; Wang, Y.C.; Tung, S.O.; Hoff, J.D.; Kotov, N. Anomalous dispersions of ‘hedgehog’ particles. Nature 2015, 517, 596–599. [Google Scholar] [CrossRef]
  27. Babu, N.J.; Nangia, A. Solubility advantage of amorphous drugs and pharmaceutical cocrystals. Cryst. Growth Des. 2011, 11, 2662–2679. [Google Scholar] [CrossRef]
  28. Ashbaugh, H.S.; Pratt, L.R. Colloquium: Scaled particle theory and the length scales of hydrophobicity. Rev. Mod. Phys. 2006, 78, 159–178. [Google Scholar] [CrossRef]
  29. Li, Z.; Loh, X.J. Water soluble polyhydroxyalkanoates: Future materials for therapeutic applications. Chem. Soc. Rev. 2015, 44, 2865–2879. [Google Scholar] [CrossRef]
  30. Schaper, L.-A.; Hock, S.J.; Herrmann, W.A.; Kühn, F.E. Synthesis and application of water-soluble NHC transition-metal complexes. Angew. Chem. Int. Ed. 2013, 52, 270–289. [Google Scholar] [CrossRef]
  31. Dill, K.A.; MacCallum, J.L. The protein-folding problem, 50 years on. Science 2012, 338, 1042–1046. [Google Scholar] [CrossRef] [Green Version]
  32. Wei, G.; Xi, W.; Nussinov, R.; Ma, B. Protein ensembles: How does nature harness thermodynamic fluctuations for life? The diverse functional roles of conformational ensembles in the cell. Chem. Rev. 2016, 116, 6516–6551. [Google Scholar] [CrossRef] [PubMed]
  33. Berne, B.J.; Weeks, J.D.; Zhou, R. Dewetting and hydrophobic interaction in physical and biological systems. Annu. Rev. Phys. Chem. 2009, 60, 85–103. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Bonn, D.; Eggers, J.; Indekeu, J.; Meunier, J.; Rolley, E. Wetting and spreading. Rev. Mod. Phys. 2009, 81, 739–805. [Google Scholar] [CrossRef]
  35. Shi, G.; Dang, Y.; Pan, T.; Liu, X.; Liu, H.; Li, S.; Zhang, L.; Zhao, H.; Li, S.; Han, J.; et al. Unexpectedly enhanced solubility of aromatic amino acids and peptides in an aqueous solution of divalent transition-metal cations. Phys. Rev. Lett. 2016, 117, 238102. [Google Scholar] [CrossRef] [PubMed]
  36. Rimola, A.; Rodríguez-Santiago, L.; Sodupe, M. Cation–π interactions and oxidative effects on Cu+ and Cu2+ binding to Phe, Tyr, Trp, and His amino acids in the gas phase. Insights from first-principles calculations. J. Phys. Chem. B 2006, 110, 24189–24199. [Google Scholar] [CrossRef]
  37. Su, P.; Li, H. Energy decomposition analysis of covalent bonds and intermolecular interactions. J. Chem. Phys. 2009, 131, 014102. [Google Scholar] [CrossRef] [Green Version]
  38. Zhang, Y.; Chen, S.; Ying, F.; Su, P.; Wu, W. Valence Bond Based Energy Decomposition Analysis Scheme and Its Application to Cation–π Interactions. J. Phys. Chem. A 2018, 122, 5886–5894. [Google Scholar] [CrossRef]
  39. Te Velde, G.; Bickelhaupt, F.M.; Baerends, E.J.; Fonseca Guerra, C.; van Gisbergen, S.J.A.; Snijders, J.G.; Ziegler, T. Chemistry with ADF. J. Comput. Chem 2001, 22, 931–967. [Google Scholar] [CrossRef]
  40. Deschamps, P.; Kulkarni, P.P.; Gautam-Basak, M.; Sarkar, B. The saga of copper(II)-L-histidine. Coord. Chem. Rev. 2005, 249, 895–909. [Google Scholar] [CrossRef]
  41. Perrin, D.D.; Sharma, V.S. Histidine complexes with some bivalent cations. J. Chem. Soc. A 1967, 724–728. [Google Scholar] [CrossRef]
  42. Sigel, H.; Griesser, R.; McCormick, D.B. On the structure of manganese (II)- and copper (II)-histidine complexes. Arch. Biochem. Biophys. 1969, 134, 217–227. [Google Scholar] [CrossRef]
  43. Mesu, J.G.; Visser, T.; Soulimani, F.; van Faassen, E.E.; de Peinder, P.; Beale, A.M.; Weckhuysen, B.M. New insights into the coordination chemistry and molecular structure of copper(II) histidine complexes in aqueous solutions. lnorg. Chem. 2006, 45, 1960–1971. [Google Scholar] [CrossRef] [Green Version]
  44. Isaac, M.; Denisov, S.A.; Roux, A.; Imbert, D.; Jonusauskas, G.; McClenaghan, N.D.; Seneque, O. Lanthanide luminescence modulation by cation-π interaction in a bioinspired scaffold: Selective detection of copper(I). Angew. Chem. Int. Ed. 2015, 54, 11453–11456. [Google Scholar] [CrossRef] [Green Version]
  45. Takeuchi, H. UV Raman markers for structural analysis of aromatic side chains in proteins. Anal. Sci. 2011, 27, 1077. [Google Scholar] [CrossRef] [Green Version]
  46. Li, S.; Hong, M. Protonation, tautomerization, and rotameric structure of histidine: A comprehensive study by magic-angle-spinning solid-state NMR. J. Am. Chem. Soc. 2011, 133, 1534–1544. [Google Scholar] [CrossRef] [Green Version]
  47. Polfer, N.C.; Oomens, J.; Moore, D.T.; von Helden, G.; Meijer, G.; Dunbar, R.C. Infrared spectroscopy of phenylalanine Ag(I) and Zn(II) complexes in the gas phase. J. Am. Chem. Soc. 2006, 128, 517–525. [Google Scholar] [CrossRef]
  48. Kruck, T.P.A.; Sarkar, B. Structure of the species in the copper (II)–L-histidine system. Can. J. Chem. 1973, 51, 3563–3571. [Google Scholar] [CrossRef]
  49. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef] [Green Version]
  50. Frisch, M.; Trucks, G.; Schlegel, H.B.; Scuseria, G.; Robb, M.; Cheeseman, J.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. Gaussian 09, Revision, A. 02; Gaussian. Inc.: Wallingford, CT, USA, 2009; p. 200. [Google Scholar]
  51. Peng, C.Y.; Ayala, P.Y.; Schlegel, H.B.; Frisch, M.J. Using redundant internal coordinates to optimize equilibrium geometries and transition states. J. Comput. Chem. 1996, 17, 49–56. [Google Scholar] [CrossRef]
  52. Rumble, J.R. CRC Handbook of Chemistry and Physics, 98th ed.; CRC Press: Boca Raton, FL, USA, 2017. [Google Scholar]
  53. Tu, Y.; Liu, H.; Shi, G.; Zhang, F.; Su, T.; Wu, Y.; Sun, J.; Zhang, L.; Zhang, S.; Fang, H. Selectivity mechanism of magnesium and calcium in cation-binding pocket structures of phosphotyrosine. Phys. Rev. E 2020, 101, 022410. [Google Scholar] [CrossRef]
  54. Hirshfeld, F.L. Bonded-atom fragments for describing molecular charge densities. Theor. Chim. Acta 1977, 44, 129–138. [Google Scholar] [CrossRef]
Figure 1. Interaction energies between His, Zn2+–His, Cu2+–His and the nearest water and the solubilities of His in ZnCl2 and CuCl2 solutions. (ac) Optimized geometric structures of the His–water, Zn2+–His–water, and Cu2+–His–water systems and the interaction energies between the imidazole rings and the nearest water molecules. (a) Optimized distance between the imidazole ring in His and the nearest water molecule is 2.3 Å, and the interaction energy between them is −3.5 kcal·mol−1. (b) Optimized distance between the imidazole ring and the nearest water molecule after Zn2+ binding decreases to 2.0 Å, and the interaction energy increases to −12.5 kcal·mol−1. (c) Optimized distance between the imidazole ring in His and the nearest water molecule after Cu2+ binding decreases to 2.0 Å, and the interaction energy increases to −13.0 kcal·mol−1. (d) Interaction energies (ΔEi) between the imidazole ring in His and the nearest water molecule, without binding to any metal ion, with Zn2+ binding, and with Cu2+ binding and the solubilities of His (SHis) in pure water, 0.4 M ZnCl2 solution, and 0.4 M CuCl2 solution. (e) Solubilities of His in CuCl2 and ZnCl2 solutions.
Figure 1. Interaction energies between His, Zn2+–His, Cu2+–His and the nearest water and the solubilities of His in ZnCl2 and CuCl2 solutions. (ac) Optimized geometric structures of the His–water, Zn2+–His–water, and Cu2+–His–water systems and the interaction energies between the imidazole rings and the nearest water molecules. (a) Optimized distance between the imidazole ring in His and the nearest water molecule is 2.3 Å, and the interaction energy between them is −3.5 kcal·mol−1. (b) Optimized distance between the imidazole ring and the nearest water molecule after Zn2+ binding decreases to 2.0 Å, and the interaction energy increases to −12.5 kcal·mol−1. (c) Optimized distance between the imidazole ring in His and the nearest water molecule after Cu2+ binding decreases to 2.0 Å, and the interaction energy increases to −13.0 kcal·mol−1. (d) Interaction energies (ΔEi) between the imidazole ring in His and the nearest water molecule, without binding to any metal ion, with Zn2+ binding, and with Cu2+ binding and the solubilities of His (SHis) in pure water, 0.4 M ZnCl2 solution, and 0.4 M CuCl2 solution. (e) Solubilities of His in CuCl2 and ZnCl2 solutions.
Ijms 23 03957 g001
Figure 2. Fluorescence and NMR spectra on His in different solutions. (a) Fluorescence spectra of His in pure water, His in ZnCl2 solution, His in CuCl2 solutions (50 mM), ZnCl2 solution and CuCl2 solution. The concentrations of His in all the solutions are 30 mg·mL−1. The concentrations of ZnCl2 and CuCl2 are 50 mM in all the samples. (b) Cross–polarization/magic angle spinning (CP/MAS) NMR C12 spectra of Cu2+–His and the control group (His in pure water). The Cδ2 and Cγ peaks are labeled.
Figure 2. Fluorescence and NMR spectra on His in different solutions. (a) Fluorescence spectra of His in pure water, His in ZnCl2 solution, His in CuCl2 solutions (50 mM), ZnCl2 solution and CuCl2 solution. The concentrations of His in all the solutions are 30 mg·mL−1. The concentrations of ZnCl2 and CuCl2 are 50 mM in all the samples. (b) Cross–polarization/magic angle spinning (CP/MAS) NMR C12 spectra of Cu2+–His and the control group (His in pure water). The Cδ2 and Cγ peaks are labeled.
Ijms 23 03957 g002
Table 1. Solubilities of His and Trp in pure water, ZnCl2 and CuCl2.
Table 1. Solubilities of His and Trp in pure water, ZnCl2 and CuCl2.
Pure WaterZnCl2CuCl2
Solubility (mg·mL−1)His41.9233.4 1244.4 1
Trp11.417.1 257.6 2
1 Salt concentration 0.4 M. 2 Salt concentration 0.5 M.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Song, Y.; Zhan, J.; Li, M.; Zhao, H.; Shi, G.; Wu, M.; Fang, H. Enhancement of the Water Affinity of Histidine by Zinc and Copper Ions. Int. J. Mol. Sci. 2022, 23, 3957. https://doi.org/10.3390/ijms23073957

AMA Style

Song Y, Zhan J, Li M, Zhao H, Shi G, Wu M, Fang H. Enhancement of the Water Affinity of Histidine by Zinc and Copper Ions. International Journal of Molecular Sciences. 2022; 23(7):3957. https://doi.org/10.3390/ijms23073957

Chicago/Turabian Style

Song, Yongshun, Jing Zhan, Minyue Li, Hongwei Zhao, Guosheng Shi, Minghong Wu, and Haiping Fang. 2022. "Enhancement of the Water Affinity of Histidine by Zinc and Copper Ions" International Journal of Molecular Sciences 23, no. 7: 3957. https://doi.org/10.3390/ijms23073957

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

Article Metrics

Back to TopTop