520 research outputs found

    A New Noncovalent Force: Comparison of P∙∙∙N Interaction with Hydrogen and Halogen Bonds

    Get PDF
    When PH(3) is paired with NH(3), the two molecules are oriented such that the P and N atoms face one another directly, without the intermediacy of a H atom. Quantum calculations indicate that this attraction is due in part to the transfer of electron density from the lone pair of the N atom to the σ(∗) antibond of a P-H covalent bond. Unlike a H-bond, the pertinent hydrogen is oriented about 180° away from, instead of toward, the N, and the N lone pair overlaps with the lobe of the P-H σ(∗) orbital that is closest to the P. In contrast to halogen bonds, there is no requirement of a σ-hole of positive electrostatic potential on the P atom, nor is it necessary for the two interacting atoms to be of differing potential. In fact, the two atoms can be identical, as the global minimum of the PH(3) homodimer has the same structure, characterized by a P···P attraction. Natural bond orbital analysis, energy decomposition, and visualization of total electron density shifts reveal other similarities and differences between the three sorts of molecular interaction

    Forty Years of Progress in the Study of the Hydrogen Bond

    Get PDF
    The author looks back at developments over the last few decades concerning the H–bond. The list of atoms involved as proton donor and acceptor has broadened dramatically, including most electronegative atoms and even metals. The factors that control the transfer of the proton across the H–bond have been elucidated and show the importance of even minor changes in its geometry. Small stretches can shut down the transfer entirely, and certain bends can force a proton to transfer against a pK gradient. Along with the recognition that a CH..O interaction can represent a true H–bond, and one with strength comparable to more traditional H–bonds, has come an understanding of its contributions to protein structure and function. The replacement of the bridging of H by any of a litany of electronegative atoms leads to similarly strong interactions, with many features virtually indistinguishable from a true H–bond. These noncovalent interactions are typically referred to as halogen, chalcogen, pnicogen, and tetrel bonds, depending upon the identity of the substitute bridging atom

    Systematic Elucidation of Factors that Influence the Strength of Tetrel Bonds

    Get PDF
    Quantum calculations are used to examine the properties of heterodimers formed by a series of tetrel-containing molecules with NH3 as universal Lewis base. TH4 was taken as a starting point, with T= C, Si, Ge, and Sn. The H atoms were replaced by various numbers of F atoms: TH3F, TF3H, and TF4 so as to monitor the effects of adding electron-withdrawing substituents. Unsubstituted TH4 molecules form the weakest tetrel bonds, only up to about 2 kcal/mol. The bond is strengthened when the H opposite NH3 is replaced by F, rising up to the 6-9 kcal/mol range. Another means of strengthening arises when the three peripheral H atoms of TH4 are replaced by F. The effect of the latter is heavily dependent on the nature of the T atom, and is particularly noticeable for larger tetrels. The two sorts of fluorination patterns are cooperative, in that their combination in TF4 yields by far the most powerful tetrel bonding agent. The tetrel bond is strengthened as the T atom moves further down the periodic table column. The strongest bond amounts to 25.5 kcal/mol for SnF4••NH3. A number of features correlate with the binding energy, but only roughly. These properties include the charge transfer, the AIM bond critical point electron density, the molecular electrostatic potential, and the stretch of the T-X covalent bond upon complex formation

    An Updated Description of the Hydrogen Bond and Related Noncovalent Bonds

    Get PDF
    The hydrogen bond is typically introduced briefly in General Chemistry as a simple electrostatic phenomenon involving a small and select group of atoms, a definition which is typically unchanged through higher levels in the curriculum. But this definition has undergone dramatic modernization of which students should be made aware. The original formulation in terms of only F, O, and N atoms has broadened very considerably, encompassing C as well as atoms from lower rows in the periodic table. The influence of hybridization, substituents, and overall charge cannot be overlooked. In addition to the Coulombic attraction, there are other “covalent” contributors such as charge transfer and polarization. Further broadening has occurred with the recognition that the bridging H can be replaced by a host of electronegative atoms in what have come to be denoted halogen, chalcogen, pnicogen, and tetrel bonds, with behavior very similar to H-bonds

    Interpretation of Spectroscopic Markers of Hydrogen Bonds

    Get PDF
    Quantum calculations are used to examine whether a AH∙∙D H-bond is unambiguously verified by a downfield shift of the bridging proton’s NMR signal or a red (or blue) shift of the AH stretching frequency in the IR spectrum. It is found that such IR band shifts will occur even if the two groups experience weak or no attractive force, or if they are drawn in so close together that their interaction is heavily repulsive. The mere presence of a proton-acceptor molecule can affect the chemical shielding of a position occupied by a proton-donor by virtue of its electron density, even if there is no H-bond present. This density-induced shielding is heavily dependent on position around the proton-acceptor atom, and varies from one group to another. Evidence of a H-bond rests on the measurement of a proton deshielding in excess of what is caused purely by the presence of the proton acceptor species

    The Hydrogen Bond: A Hundred Years and Counting

    Get PDF
    Since its original inception, a great deal has been learned about the nature, properties, and applications of the H-bond. This review summarizes some of the unexpected paths that inquiry into this phenomenon has taken researchers. The transfer of the bridging proton from one molecule to another can occur not only in the ground electronic state, but in various excited states as well. Study of the latter process has developed insights into the relationships between the nature of the state, the strength of the H-bond, and the height of the transfer barrier. The enormous broadening of the range of atoms that can act as both proton donor and acceptor has led to the concept of the CH···O HB, whose properties are of immense importance in biomolecular structure and function. The idea that the central bridging proton can be replaced by any of various electronegative atoms has fostered the rapidly growing exploration of related noncovalent bonds that include halogen, chalcogen, pnicogen, and tetrel bonds

    On the Capability of Metal-Halogen Groups to Participate in Halogen Bonds

    Get PDF
    A number of halogen (X) atoms were covalently attached to a metal (M) and the ability of the X atom to act as electron acceptor in a halogen bond to nucleophile NCH was assessed. Both Cl and Br were considered as halogen atom, with NH3 and CO as other ligands attached to the metal. Metals tested were Ti, Mn, and Zn in various combinations of oxidation state, coordination, and overall charge. In the majority of cases, the strong electron-releasing power of the metal imbues the halogen atom with a high negative partial charge and minimizes the development of a σ-hole. As such, the M atom is generally a stronger attractor for the incoming nucleophile than is the halogen. Nonetheless, there are cases where a halogen bond can form such as Ti(CO)4Br+, TiCl3+, and MnCl4+, each with a different coordination. A requisite of halogen bond formation is generally an overall positive charge, although neutral species can engage in such bonds, albeit much weaker

    Differential Binding of Tetrel-Bonding Bipodal Receptors to Monatomic and Polyatomic Anions

    Get PDF
    Previous work has demonstrated that a bidentate receptor containing a pair of Sn atoms can engage in very strong interactions with halide ions via tetrel bonds. The question that is addressed here concerns the possibility that a receptor of this type might be designed that would preferentially bind a polyatomic over a monatomic anion since the former might better span the distance between the two Sn atoms. The binding of Cl− was thus compared to that of HCOO−, HSO4−, and H2PO4− with a wide variety of bidentate receptors. A pair of SnFH2 groups, as strong tetrel-binding agents, were first added to a phenyl ring in ortho, meta, and para arrangements. These same groups were also added in 1,3 and 1,4 positions of an aliphatic cyclohexyl ring. The tetrel-bonding groups were placed at the termini of (-C≡C-)n (n = 1,2) extending arms so as to further separate the two Sn atoms. Finally, the Sn atoms were incorporated directly into an eight-membered ring, rather than as appendages. The ordering of the binding energetics follows the HCO2− \u3e Cl− \u3e H2PO4− \u3e HSO4− general pattern, with some variations in selected systems. The tetrel bonding is strong enough that in most cases, it engenders internal deformations within the receptors that allow them to engage in bidentate bonding, even for the monatomic chloride, which mutes any effects of a long Sn···Sn distance within the receptor

    Comparison of Bifurcated Halogen with Hydrogen Bonds

    Get PDF
    Bifurcated halogen bonds are constructed with FBr and FI as Lewis acids, paired with NH3 and NCH bases. The first type considered places two bases together with a single acid, while the reverse case of two acids sharing a single base constitutes the second type. These bifurcated systems are compared with the analogous H-bonds wherein FH serves as the acid. In most cases, a bifurcated system is energetically inferior to a single linear bond. There is a larger energetic cost to forcing the single σ-hole of an acid to interact with a pair of bases, than the other way around where two acids engage with the lone pair of a single base. In comparison to FBr and FI, the H-bonding FH acid is better able to participate in a bifurcated sharing with two bases. This behavior is traced to the properties of the monomers, in particular the specific shape of the molecular electrostatic potential, the anisotropy of the orbitals of the acid and base that interact directly with one another, and the angular extent of the total electron density of the two molecules

    Assessing the Possibility and Properties of Types I and II Chalcogen Bonds

    Get PDF
    Type I and II halogen bonds are well-recognized motifs that commonly occur within crystals. Quantum calculations are applied to examine whether such geometries might occur in their closely related chalcogen bond cousins. Homodimers are constructed of the R1R2C=Y and R1R2Y monomers, wherein Y represents a chalcogen atom, S, Se, or Te; R1 and R2 refer to either H or F. A Type II (T2) geometry wherein the lone pair of one Y is closely aligned with a σ-hole of its partner represents a stable arrangement for all except YH2, although not all such structures are true minima. The symmetric T1 geometry in which each Y atom serves as both electron donor and acceptor in the chalcogen bond is slightly higher in energy for R1R2C=Y, but the reverse is true for R1R2Y. Due to their deeper σ-holes, the latter molecules engage in stronger chalcogen bonds than do the former, with the exception of H2Y, whose dimers are barely bound. The interaction energies rise as the Y atom grows larger: S \u3c Se \u3c Te
    corecore