Structure and stretching dynamics of water molecules around an amphiphilic amide from FPMD simulations: A case study of N,N-dimethylformamide

https://doi.org/10.1016/j.molliq.2020.112524Get rights and content

Highlights

  • Dynamics and spectral properties of water molecules in the vicinity of an amphiphilic amide, N,N-dimethylformamide

  • The carbonyl-water hydrogen bonds are stronger than water-water hydrogen bonds.

  • The methyl substituents on nitrogen of DMF impose weak hydrophobic interaction.

  • The hydrophilic carbonyl oxygen forms a strong hydrogen bond with the neighboring water.

  • The heterogeneity behavior of groups present in DMF gives rise different spectral signature of OH modes of water molecule.

Abstract

N,N-Dimethylformamide (DMF) is a unique tertiary amphiphilic amide, where the presence of a hydrophilic aldehyde group favors hydrogen bond acceptance, but two hydrophobic methyl substituents inhibit interaction with water molecules. As a result, the water molecules encounter two different environments in the vicinity of DMF molecule: around the hydrophobic nitrogen site; and near the hydrophilic carbonyl oxygen. We employ first principles molecular dynamics methods to simulate an aqueous solution of DMF using PBE functional with Grimme's D3 dispersion correction at 330 K. Investigations on the liquid structure to understand inter-atomic interactions in amides attract much attention. We calculated various structural and dynamical properties along with vibrational stretching frequency of water molecules to understand heterogeneously affected water molecules by an amphiphilic amide molecule. In solvated DMF, the first peak minimum of the N-OW and OC-OW radial distribution functions (‘w’ subscript denotes atoms of water) are located at 5.72 and 3.16 Å, respectively. These distance cutoffs decide the boundary of the solvation shell. At OC-HW distance 2.45 Å, the deep peak minimum indicates stable OC … HW hydrogen bond. Previous Monte Carlo simulations reported the presence of hydrogen bonds between the oxygen site of DMF and hydrogen of water. The time-series wavelet method was used to compute the time-dependent frequencies of the hydroxyl groups of water. The average frequency of the OH modes inside the Cdouble bondO solvation shell (~3364 cm1) in DMF is higher than bulk (~3337 cm1), and the trend matches with an aqueous solution of acetone. In nitrogen hydration shell, the intense band resembles the bulk frequency distribution, and a narrow distinctive peak at the high-frequency side (range ~3650–3750 cm1) represents the non‑hydrogen bonded or dangling OH groups. Raman spectroscopy in hydrophobic TBA and air/water interface displayed dangling OH stretch peak at ~3660 cm1 and ~3710 cm1, respectively. Our calculation of the frequency distribution, frequency-frequency correlation function, and hydrogen bond dynamics show the water molecules at bulk behave as in pure water. Inside the solvation shell of aminic nitrogen, non‑hydrogen bonded OH modes with a dangling lifetime ~0.38 ps dominates over the water-water hydrogen bonds. The time-dependent decay of the frequency correlation inside Cdouble bondO solvation shell has three decay components, a rapid decay, then, an intermediate component (~1.52 ps) corresponding to the lifetime of the carbonyl-water hydrogen bond, and the longer timescale ~11.97 ps representing the residence time of water molecules in the vicinity of carbonyl oxygen. We find that the carbonyl-water hydrogen bond (OC … HW) is stronger than the water-water hydrogen bond. The methyl substituents on the nitrogen of DMF impose weak hydrophobic interaction, and the hydrophilic carbonyl oxygen forms a strong hydrogen bond with the neighboring water resulting in localized dynamics.

Introduction

In peptides, the basic structural repeating unit is the carbonyl carbon to nitrogen bond, commonly known as the amide linkage. Amides are carboxylic acid derivatives, where an amine replaces the hydroxyl group, and the resulting peptide bonds formed via covalent linkages in the polypeptides confer structural rigidity and serve as structural materials in aqueous biochemical processes preventing hydrolysis [[1], [2], [3], [4], [5]]. The structure, dynamics, and overall bonding of aqueous solutions of these polyamides depend on the strength of the amide-water hydrogen bonding interactions. To study amide facilitated changes in structural and stretching dynamics of water molecules and the solvent effect on the amide moieties, we investigate water interactions with an exceptional tertiary amide DMF, consisting of two hydrophilic amino (single bondNH2) and aldehyde (single bondCHO) sites along with hydrophobic methyl groups. Moreover, it has been reported recently that the hydrophobic solubility may facilitate the stability of the protein in an aqueous environment [6]. DMF can act as both a proton donor and an acceptor. DMF can only be a proton acceptor (or H-bond acceptor) on substitution of the two hydrogen atoms of the amino group by methyl functionality in the simplest amide molecule formamide and it forms very weak Csingle bondH…OW hydrogen bond [1]. Dialkyl substitution hinders Nsingle bondH…Odouble bondC hydrogen bonding and makes DMF unique 3° amide to be studied. DMF is a common solvent in chemical reactions due to its low evaporation rate, and high dielectric constant [[7], [8], [9], [10]]. The dielectric constants are varied to make the aqueous mixtures useful solvents. On hydrating the DMF molecule, it undergoes structural changes, the behavior of water molecules in DMF-water mixtures vary with the surrounding solvent medium. In DMF solvation, water molecules form hydrogen bonds, and the hydrophobic interactions of the methyl functional groups affect the vicinal water molecules. The presence of a diverse environment in DMF attracts our attention to explore and comment on the structure, dynamics, and hydrogen bonding patterns prevailing in amide-water mixtures. Previously several experimental [[11], [12], [13], [14]] and theoretical studies [[15], [16], [17], [18], [19], [20], [21], [22], [23], [24]] on pure and aqueous DMF solutions were reported. The time-domain reflectometry (TDR) method was used to study the dielectric properties of DMF-water mixtures to understand hydrogen bonding and intramolecular rotations [25]. Molecular dynamics (MD) simulations using the OPLS all-atom force field showed the composition dependence and the clustering of water molecules in an aqueous solution of DMF [1]. It was seen that dipolar interactions existed between DMF and the solvent in pure solution, but hydrogen-bonded associations were not detected in the measurement of NMR chemical shift [26]. The calculation of the enthalpy of solution did not report the hydrophobic effect in mixtures of tetraalkylammonium ions [27]. In the DMF-water system, the physical properties like densities, viscosities, and the dynamical properties depicting molecular mobility like the diffusion coefficient vary as a function of solute composition and temperature variation [8,9]. These physicochemical properties [8,9,[28], [29], [30]] arise from the interactions of a single DMF molecule with its well-defined solvent environment, and the hydrogen bonds between water molecules and DMF result in the formation of water clusters. Two-dimensional correlation spectroscopy [31] found out the composition-dependent spectral shifts in the specific stretching bands. With the increase of the water content, two different regions were identified due to the appearance of Csingle bondH and Osingle bondH bands. [31] Numerous investigations [21,[32], [33], [34], [35], [36], [37], [38]] on the microscopic environment in mixtures of water and organic liquids revealed the interesting properties in the liquid structure arise from the water hydrogen-bonded structures [39]. Reorientation time correlation is a sensitive reporter of the intermolecular interactions of water molecules in amide-water mixtures [20,29,[40], [41], [42], [43], [44], [45]]. The measurement of the NMR spin-lattice relaxation times at 5, 25, and 45 °C depicted the hydrophilic and hydrophobic interactions between amide and water molecules and its impact on molecular reorientational dynamics [29]. Molecular dynamics study of aqueous amide mixtures addressed the relationship between concentration fluctuations and microheterogeneity [39]. Monte Carlo simulations [10] of dilute amide-water solutions performed at ambient conditions reported hydrogen bonds between the carbonyl oxygen of amide and the water molecules, but the solvent molecules around hydrophobic methyl showed little interaction with the solutes [10]. The other study [16] on the water-DMF interaction used the same simulation protocol and characterized Cdouble bondO…HW hydrogen bonds between the carboxyl oxygen site of DMF and water hydrogen by the calculations of the radial distribution functions. Statistical perturbation theory [16] was employed to monitor the stabilization of the DMF molecule, and it was observed that the energy changes involved in the hydration process stabilized DMF in its planar configuration. In this paper, we present the structural details that help in understanding the solvent structure around DMF, the phenomena of vibrational spectral diffusion of water molecules, and the strength and lifetime of hydrogen bonds affecting the dynamics in the amide-water systems.

Section snippets

Computational methods

First principles molecular dynamics (FPMD) [46] simulations of 50 water molecules and single DMF molecule were carried out in a cubic box with box vector 11.71 Å, using the Quickstep module [47] implemented in the CP2K [48] software package. The box length was calculated after performing the simulations within NpT ensemble. The dual centered Gaussian-type orbitals [47,49] and plane waves available within the Gaussian plane wave approach solve the self-consistent Kohn-Sham equations of density

Solvation shell structure in aqueous DMF

Structural detailing in liquid amides has drawn considerable interest in physical chemists. X-ray diffraction measurements determined the liquid structure of DMF [58], and later the structure was compared with the molecular structure of formamide (FA), declaring the absence of associated species in liquid DMF [59]. Understanding the atom-atom intermolecular interactions required support from reliable experimental and simulated radial distribution functions. To elucidate the intermolecular

Discussion

In this section, we compare the structural aspects of the amide solvation shell, spectral properties, and the dynamical response of the surrounding water molecules in N,N-dimethylformamide (DMF) with the earlier reported other aqueous amides studied using similar methods. Car-Parrinello molecular dynamics (CPMD) simulations with a single formamide (FA) dissolved in 31 solvent molecules, followed by the time-series wavelet analysis, calculated the time-dependent vibrational stretch frequencies. [

Conclusions

An aqueous solution of DMF was studied using first principles molecular dynamics simulation using the Perdew-Burke-Ernzerhof (PBE) [51,52] exchange-correlational functional and Grimme's D3 [55,56] dispersion correction using TZV2P basis set at a temperature of 330 K. The hydrophilic aldehyde carbonyl participates in hydrogen bond formation with the nearest neighboring water molecules. The DMF-water hydrogen bond between the single bondCHO group and the corresponding hydrogen or oxygen atom of the adjacent

CRediT authorship contribution statement

Aritri Biswas: Formal analysis, Visualization, Validation, Writing - original draft. Bhabani S. Mallik: Investigation, Conceptualization, Methodology, Writing - review & editing, Supervision, Project administration, Funding acquisition.

Acknowledgments

The authors acknowledge financial support (EMR/2016/004965) for this work from the Department of Science and Technology, India. Aritri Biswas likes to thank the Ministry of Human and Resources Development, India, for the Ph.D. fellowship.

References (107)

  • E. Siebel et al.

    Solid State Ionics

    (1997)
  • H.N. Dilip et al.

    J. Mol. Liq.

    (2019)
  • S. Yashonath et al.

    Chem. Phys.

    (1991)
  • M. Cordeiro et al.

    J. Mol. Struct.-Theochem - J MOL STRUC-THEOCHEM

    (2006)
  • M. Chalaris et al.

    J. Mol. Liq.

    (2002)
  • M. Nakahara

    C. Wakai

  • J. VandeVondele et al.

    Comput. Phys. Commun.

    (2005)
  • Y.P. Puhovski et al.

    J. Mol. Liq.

    (2003)
  • S. Biswas et al.

    J. Mol. Liq.

    (2015)
  • J.B. Asbury et al.

    Chem. Phys. Lett.

    (2003)
  • Y. Lei et al.

    J. Phys. Chem. A

    (2003)
  • M.A.M. Cordeiro et al.

    J. Braz. Chem. Soc.

    (2004)
  • R.A. Musah et al.

    J. Am. Chem. Soc.

    (1997)
  • M.S. Butler

    J. Nat. Prod.

    (2004)
  • M.M. Heravi et al.

    RSC Adv.

    (2018)
  • G. Guarino et al.

    J. Chem. Eng. Data

    (1985)
  • C.D. Volpe et al.

    J. Chem. Eng. Data

    (1986)
  • W.L. Jorgensen et al.

    J. Am. Chem. Soc.

    (1985)
  • R. Konrat et al.

    J. Phys. Chem.

    (1990)
  • M. Holz et al.

    J. Chem. Phys.

    (1996)
  • H. Borrmann et al.

    J. Chem. Soc. Perkin Trans.

    (2000)
  • A. Panuszko et al.

    J. Phys. Chem. B

    (2008)
  • M. Chalaris et al.

    J. Chem. Phys.

    (2000)
  • J.M.M. Cordeiro et al.

    Z. Für Naturforschung A

    (2014)
  • J.W. Essex et al.

    J. Phys. Chem.

    (1995)
  • J. Zielkiewicz

    Phys. Chem. Chem. Phys.

    (2000)
  • J. Gao

    J. Am. Chem. Soc.

    (1993)
  • J. Gao et al.

    J. Phys. Chem.

    (1996)
  • G.R. Desiraju

    Acc. Chem. Res.

    (1991)
  • A.C. Kumbharkhane et al.

    J. Solut. Chem.

    (1993)
  • A.D. Buckingham

    Can. J. Chem.

    (1960)
  • C. Visser et al.

    Recl. Trav. Chim. Pays-Bas

    (2010)
  • J.M. Bernal-García et al.

    J. Chem. Eng. Data

    (2008)
  • M. Okada et al.

    Bull. Chem. Soc. Jpn.

    (2012)
  • S. Biswas et al.

    J. Chem. Eng. Data

    (2014)
  • Z. Xu et al.

    J. Chem. Phys.

    (2006)
  • S. Dixit et al.

    Nature

    (2002)
  • L. Dougan et al.

    J. Chem. Phys.

    (2004)
  • S.K. Allison et al.

    Phys. Rev. B

    (2005)
  • S. Weerasinghe et al.

    J. Chem. Phys.

    (2003)
  • M.E. Lee et al.

    J. Chem. Phys.

    (2005)
  • O. Acevedo et al.

    Wiley Interdiscip. Rev. Comput. Mol. Sci.

    (2014)
  • A. Perera et al.

    J. Chem. Phys.

    (2004)
  • L. Zoranić et al.

    J. Chem. Phys.

    (2009)
  • H. Weingärtner et al.

    J. Solut. Chem.

    (1978)
  • E.S. Baker et al.

    J. Phys. Chem.

    (1985)
  • M. Nakahara et al.

    J. Phys. Chem.

    (1995)
  • M. Ueno et al.

    J. Solut. Chem.

    (2004)
  • K. Yoshida et al.

    J. Chem. Phys.

    (1998)
  • R. Car et al.
  • Cited by (20)

    • Molecular self-assembling structure and formation mechanism of lubricating grease: A computational simulation approach

      2023, Tribology International
      Citation Excerpt :

      The radial distribution function (RDF) is a method of statistical mechanics that reflects the microstructure of materials. It describes the ratio of the probability density of the occurrence of another molecule or atom within a radius r around a molecule or atom relative to the probability density of a random distribution, and is often used to reflect the degree of order of the microstructure of soft matter materials, and thus reflecting the stability of the microstructure [53,54]. Fig. 10 shows the RDF (g(r)) for the inter-carboxylate interactions of Li-ST(a) and Li-12HS(b) molecules at different temperatures during the MD process, where the first sharp peaks extending beyond the range of the graph are the intramolecular peaks of O-Li+, and the other peaks correspond to the intermolecular peaks of O and Li+.

    • Vibrational spectral dynamics and ion-probe interactions of the hydrogen-bonded liquids in 1-ethyl-3-methylimidazolium bis(trifluoromethylsulfonyl)imide

      2022, Chemical Physics
      Citation Excerpt :

      Moreover, wavelet analysis was applied repeatedly through each probe's entire 1 ns trajectory to calculate the vibrational stretching frequencies of all the OD bonds at each timeframe. Further, our previous reports have presented an in-depth analysis of the wavelet method [17,74-78]. We determine the O-D stretch spectral signature of each vibrational probe molecule in the normalized frequency distribution from the wavelet-derived instantaneous fluctuating frequencies [70,72].

    • Structure and rotational dynamics of water around hydrogen peroxide

      2022, Journal of Molecular Liquids
      Citation Excerpt :

      Water tends to exhibit different structural, dynamical, and spectral properties in the different regions of the above-discussed solute systems. From the previous studies concerning purely hydrophobic solutes like tetramethylammonium cation, partly hydrophobic solutes like N, N-dimethylformamide [48], and solutes with no hydrophobic hindrance like amide [49], significantly different RDFs of hydrogen and oxygen of water molecules concerning the different atomic centers of the concerned solute were observed. N, N-dimethylformamide [48] has a hydrophilic aldehyde (–CHO) site and hydrophobic methyl (–CH3) group around which water O can act as proton donor and acceptor.

    • Revisiting OD-stretching dynamics of methanol‑d<inf>4</inf>, ethanol-d<inf>6</inf> and dilute HOD/H<inf>2</inf>O mixture with predefined potentials and wavelet transform spectra

      2022, Chemical Physics
      Citation Excerpt :

      Shorter and stronger hydrogen bonds redshift the vibrational frequency, and the weaker hydrogen bonds shift the spectral band towards the blue side of the spectrum. The continuous fluctuation of the hydrogen-bonding networks’ structural evolution with time affects the probe’s vibrational frequency, referred to as spectral diffusion [8,11–15]. The infrared spectroscopic properties of the fundamental hydroxyl stretch in saturated alcohols attract immense interest.

    View all citing articles on Scopus
    View full text