Next Article in Journal / Special Issue
An Electrochemical Impedance Study of Alkaline Water Splitting Using Fe Doped NiO Nanosheets
Previous Article in Journal / Special Issue
The Importance of Interphases in Energy Storage Devices: Methods and Strategies to Investigate and Control Interfacial Processes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Vibrational Analysis of Benziodoxoles and Benziodazolotetrazoles

Department of Chemistry, Southern Methodist University, 3215 Daniel Avenue, Dallas, TX 75275, USA
*
Author to whom correspondence should be addressed.
Submission received: 27 March 2021 / Revised: 27 April 2021 / Accepted: 5 May 2021 / Published: 14 May 2021
(This article belongs to the Special Issue Physical Chemistry Perspectives for the New Decade)

Abstract

:
Tetrazoles are well known for their high positive enthalpy of formation which makes them attractive as propellants, explosives, and energetic materials. As a step towards a deeper understanding of the stability of benziodazolotetrazole (BIAT)-based materials compared to their benziodoxole (BIO) counterparts, we investigated in this work electronic structure features and bonding properties of two monovalent iodine precursors: 2-iodobenzoic acid and 5-(2-iodophenyl)tetrazole and eight hypervalent iodine (III) compounds: I-hydroxybenzidoxolone, I-methoxybenziodoxolone, I-ethoxybenziodoxolone, I-iso-propoxybenziodoxolone and the corresponding I-hydroxyben ziodazolotetrazole, I-methoxybenziodazolotetrazole, I-ethoxybenziodazolotetrazole and I-iso- propoxybenziodazolotetrazole. As an efficient tool for the interpretation of the experimental IR spectra and for the quantitative assessment of the I−C, I−N, and I−O bond strengths in these compounds reflecting substituent effects, we used the local vibrational mode analysis, originally introduced by Konkoli and Cremer, complemented by electron density and natural bond orbital analyses. Based on the hypothesis that stronger bonds correlate with increased stability, we predict that, for both series, i.e., substituted benziodoxoles and benziodazolotetrazoles, the stability increases as follows: I-iso-propoxy < I-ethoxy < I-methoxy < I-hydroxy. In particular, the I−N bonds in the benziodazolotetrazoles could be identified as the so-called trigger bonds being responsible for the initiation of explosive decomposition in benziodazolotetrazoles. The new insight gained by this work will allow for the design of new benziodazolotetrazole materials with controlled performance or stability based on the modulation of the iodine bonds with its three ligands. The local mode analysis can serve as an effective tool to monitor the bond strengths, in particular to identify potential trigger bonds. We hope that this article will foster future collaboration between the experimental and computational community being engaged in vibrational spectroscopy.

Graphical Abstract

1. Introduction

Organic polyvalent iodine (III) and (V) reagents have gained popularity as oxidants, electrophiles, and radical sources in various synthetically useful organic transformations [1,2,3,4,5,6,7,8], as well as in organometallics [9], materials, and polymer chemistries [10]. These compounds contain an iodine atom which bonds to a carbon atom via conventional 2-center-2-electron (2c-2e) covalent bond, but also to two or more ligands (L) via 3-center-4-electron (3c-4e) bonds [11,12], dubbed hypervalent (HV) [13]. According to this model, only p-orbitals of the central atom are involved in the formation of the 3c-4e bonds. Since publication of the aforementioned pioneering works, a number of additional studies have appeared in the literature discussing in more detail the nature of HV bonds [14,15,16,17]. The presence of weak HV bonds is largely responsible for the characteristic structural features of HV iodine (HVI) molecules (e.g., presence of the near-linear L−I−L fragment(s)), as well as for their stability and reactivity [18,19]. Donor atoms of the ligands must be electronegative (e.g., F, O, N, or Cl) to promote stability. Among said ligands, carboxylates are the most widely studied and HVI(III) compounds such as (dicarboxyloxy)iodo arenes (ArI(O2CR)2) [20] with linear arrangements of O atoms from the ligands and the central iodine atom) and benziodoxoles (where the carboxylate ligand and the HVI atom form a ring) [8,21] have found numerous applications. Tetrazolates (RCN4) resemble carboxylates (RCO2) in many respects (e.g., basicity, nucleophilicity, and complexation ability [22]), which prompted Tsarevsky et al. to investigate the propensity of tetrazole (TZ) for bonding to iodine(III) atoms. Several examples of TZ-containing HVI(III) compounds were synthesized and studied, including acyclic [23], pseudocyclic (with iodonium-like structure) [24], and heterocyclic benziodoazotetrazole (BIAT) derivatives [25]. Each of these compounds, like their carboxylate analogues, are strong oxidants. However, unlike the caboxylates, when heated BIATs decompose vigorously and even explosively. TZs are well known for their high positive enthalpy of formation ( Δ H f ) [26,27,28,29], which makes them attractive as propellants, explosives, and energetic materials [30], but HV compounds with one or more tetrazolate ligands are markedly more unstable than the parent free ligands [23,24].
An important question has been raised: which molecular parameters make TZs so explosive, and, even more importantly, how can their thermal stability be controlled. To answer this question, Tsarevsky et al. determined the enthalpies of thermal decomposition ( Δ H d ) for BIAT derivatives with axial Cl, OH, OMe, and OAc ligands. However, they did not find any significant correlation between Δ H d and structural parameters (e.g., lengths of either of the HV bonds or L−I−N valence angles), nucleophilicity, or field electronic parameters of the ligand [25]. It has been suggested that crystal packing in the solid state, i.e., the shearing force between packing layers may well account for the (in)stability of TZs, influencing both temperatures of decomposition and Δ H d , which complicates any direct comparison between Δ H d and structural parameters [31,32,33]. A common way to answer this question from a theoretical perspective is to use bond dissociation energies (BDE) [34,35] as a tool to identify the so-called trigger bond, which is first to break and therefore assesses the stability of a material [36]. However, a caveat is needed. BDE is not a suitable way to measure the intrinsic strength of a chemical bond because it also contains geometry relaxation and electronic structure reorganization of the dissociated fragments [37,38,39,40,41,42,43]. Therefore, this work refers to the local vibrational mode (LMV) theory originally proposed by Konkoli and Cremer [44,45,46,47,48], which provides a quantitative measure of intrinsic bond strength based on vibrational spectroscopy [49].
As a step towards understanding more about the nature and stability of TZ-based energetic materials from an electronic structure perspective, this work investigates the iodine-containing bond strengths in two precursor monovalent iodine compounds, 2-Iodobenzoic acid (IBA) and 5-(2-Iodophenyl)tetrazole (IPT), as well as in HVI(III) compounds derived thereof, namely substituted BIOs 14 and BIATs 58 (shown in Figure 1), utilizing local vibrational force constants derived from LVM. This work is complemented with electron density information and natural bond orbitals (NBOs) to assess and predict chemical bond properties which correlate with explosive capabilities, and to explore the suggestion that a strongly positive molecular center (i.e., iodine) may serve as a reliable indicator for sensitive materials [50,51,52]. In addition, this work applies another feature of the LMV theory: the comprehensive analysis of experimental (exp) IR spectra [46,48]. An overarching aim of this research is to inspire the experimental and theoretical chemistry communities working in the field of vibrational spectroscopy to foster future collaborations.

2. Computational and Experimental Methods

To validate the model chemistry used in this work, experimental geometry coordinates and IR frequencies of BIO-1 (see Table 1 and Table 2) were utilized as benchmarks for several levels of theory. Table 1 shows comparison between experimental (exp) and theor geometry parameters, and Table 2 summarizes the statistical analysis of exp IR frequencies ( ω ( e x p ) ) and theor normal mode frequencies ( ω μ ) of BIO-1; calculated using 13 levels of theory. The B3LYP [53,54,55,56], B3LYP-D3 [57], B3LYP-D3BJ [58], M06-2X [59], and ω B97X-D [60] density functionals were combined with the following basis sets: aug-cc-pVTZ-[PP for iodine] (acpVTZ-PP) [61,62,63,64,65,66], Def2-TZVP (TZVP) [67,68], Def2-TZVPP (TZVPP) [69,70]. The Def2-TZVPD basis set was also included, but only in combination with M06-2X because this level of theory was recently applied successfully for a computational study of phenomenal molecular systems containing λ 3 -iodanes acting as biaxial halogen bond (XB) donors [71]. For the theor vibrational frequencies, the following scaling factors were applied to correct for the harmonic approximation of the Morse potential [72,73,74,75,76,77,78,79]: 0.968 (B3LYP/acpVTZ-PP), 0.965 (B3LYP/TZVP), 0.963 (B3LYP/TZVPP), 0.956 (M06-2X/acpVTZ-PP), 0.946 (M06-2X/TZVP), 0.957 ( ω B97X-D/acpVTZ-PP), and 0.955 ( ω B97X-D/TZVP) [73,74,75,76,77,78,79].
The results of the first attempts to determine the molecular structure of I-hydroxyben ziodoxolone (BIO-1) were published in 1964 by Shefter and Wolf [80], who noted that the endocyclic I−O bond was markedly longer (2.30 Å) than the sum of the covalent radii of the two atoms, which was attributed to steric strain or greater ionic character compared to the normal covalent I−O bond. A more detailed analysis, including all bond lengths and valence angles in the molecule, was provided a year later by the same authors [81]. Unfortunately, the published structure is imprecise (which becomes apparent by simple inspection of the clearly nonplanar benzene ring) and no further structural data have been reported since the publication of the two aforementioned papers. We therefore prepared crystals of BIO-1 (by recrystallization from water) and were able to obtain a more accurate structure (Figure 2). The complete experimental details and structural data are provided in the Supporting Materials. This new structure was used for the analyses and comparison with the computed structural parameters, which allowed for proper selection of the computational method to be employed.
Calculations performed at the ω B97X-D/Def2-TZVP level of theory were in closest agreement with exp IR frequencies (average % error = 1.50, MAE = 12.99, standard deviation = 16.25, RMSE = 20.48), whereas comparison of exp and theor geometry parameters is less conclusive. Therefore, said model chemistry was utilized for the remainder of this work. All DFT calculations were performed with tight convergence criteria and superfine integration grids [82], and all stationary points were confirmed to be minima by the absence of imaginary normal mode frequencies.
In addition to serving as a popular analytical tool, modern vibrational spectroscopy [83,84,85] is an excellent source of information on the electronic structure a molecule. Specifically, a new quantitative measure for the intrinsic strength of a chemical bond is derived from vibrational force constants. However, one has to consider that normal vibrational modes are generally delocalized over the molecule due to coupling of the atomic motions [86,87,88]. Therefore, one cannot directly derive an intrinsic measure of bond strength from said normal modes, a fact that has already been discussed since the early days of vibrational spectroscopy. (The reader may refer to review articles, e.g., by Quack [89], Halonen [90], and Jensen [91] for a more in-depth discussion and to Ref. [49] for a historical overview.) Furthermore, it can be difficult to assign a peak on an experimental spectrum to a normal mode involving the vibrations of a particular functional group, particularly in the mid and lower frequency domains. Both problems are addressed by Local Vibrational Mode (LVM) theory, originally introduced by Konkoli and Cremer [44,45,46,47,48]. In Ref. [49], a comprehensive discussion of the underlying theory of LVMs is provided; therefore, in the following, only some essential features are summarized.
As outlined in Ref. [49], LVMs have a number of unique properties. Zou, Kraka, and Cremer [46,47] verified the uniqueness of the LVMs via an adiabatic connection scheme between local and normal vibrational modes. In contrast to normal mode force constants, LVM force constants have the advantage of not being dependent on the choice of coordinates used to describe the target molecule. The latter are highly sensitive to electronic structure differences (e.g., caused by changing a substituent) and directly reflect the intrinsic strength of a bond or weak chemical interaction, as has been shown by Zou and Cremer [92]. Thus, LVM stretching force constants have been utilized as a unique measure of the intrinsic strength of chemical bonds [38,39,48,93,94,95,96,97,98,99,100,101,102,103,104] and weak chemical interactions [19,40,105,106,107,108,109,110,111,112,113,114,115,116,117,118,119,120,121,122,123,124,125,126] based on vibrational spectroscopy. LVM theory has previously been applied for the description of bonding in a diverse set of 34 HVI compounds, the majority of which are λ 3 -iodanes [19]. In this work, LVM theory and associated LVM stretching force constants were applied to investigate the I−C, I−N, and I−O bonds composing the ‘T’-shaped molecular frameworks in a set of four BIATs and four BIOs.
In addition to local force constants and frequencies, the LVM analysis has led to a new way of analyzing vibrational spectra. Any normal vibrational mode l μ can be decomposed into LVM contributions [46,48], leading to detailed analysis of vibrational spectra and a wealth of information about structure and bonding [97,107]. Decomposition of normal modes into percent LVM contributions, also called Characterization of Normal Modes (CNM), calculates for each normal mode l μ the overlap S n μ with each local mode vector a n x , with both vectors given in Cartesian coordinates x , according to [46,48]:
S n μ = ( a n x , l μ ) 2 ( a n x , a n x ) ( l μ , l μ ) ; w i t h a n x = L a n
( a , b ) in Equation (1) is the short notation for the scalar product of two vectors of a and b
( a , b ) = i , j a i O i j b j
O i j is an element of the metric matrix O . We generally use the force constant matrix f x as metric, namely O = f x to account for the influence of the electronic structure [46,48]. This leads to the contribution C n μ of local mode a n to the normal mode l μ given by:
C n μ = S n μ m N v i b S m μ
i.e., a completely localized normal mode l μ has a C n μ value of 1 (corresponding to 100% if C n μ is given as a percentage). The CNM procedure was recently applied to assess the usefulness of Vibrational Stark Effect probes [102], and to analyze π -hole interactions between aryl (Ar) donors and small molecule acceptors [126]. In this work, CNM is used to interpret exp and theor spectra.
DFT calculations were carried out with GAUSSIAN16 [127] and the standard frequency calculations were followed by LVM analysis, performed with the LModeA program [49,128]. The spectroscopic analysis was complemented with NBO populations calculated with the NBO6 package [129,130,131]. Electron densities ( ρ ( r b ) ) and energy densities ( H ( r b ) ) at the I−C, I−O, and I−N bond critical points r b determined with the AIMAll package [132]. The covalent nature of the bonds was assessed following the Cremer–Kraka criterion, which implies that covalent bonding is characterized by a negative energy density, whereas electrostatic interactions are indicated by positive energy density values [133,134,135].
Commercially available IBA (98% purity) was purchased from Sigma-Aldrich, St Louis. IPT, BIO-1 [25,136], and BIAT-5 [25] were prepared using previously reported procedures. The identity and purity of synthesized compounds were confirmed by nuclear magnetic resonance (NMR) spectroscopy. Infrared (IR) spectra were collected on a Nicolet iS5 IR spectrometer with iD7 attenuated total reflectance (ATR) accessory and diamond crystal, and were formatted using OriginPro 9.1.0 (64 bit) [137]. Data were recorded in the domain of 4000–400 cm 1 . The number of scans were set to 16 with spectral resolution of 4 cm 1 . To ensure no contamination during the analysis, a background spectrum was collected every time before collecting the sample spectrum. All experiments were performed under ambient conditions.

3. Results/Discussion

3.1. Experimental and Theoretical IR-Spectra

Figure 3a,b show the exp and theor IR spectra of IBA, respectively. All exp compounds were in the solid phase under standard conditions, whereas theor normal mode vibrational frequencies were calculated in the gas phase. This difference in phase is responsible for the sharper, more well-defined peaks of the theor IR spectra. Additionally, corresponding peaks of the theor spectra have higher frequencies compared to the exp spectra which is also attributable to the phase difference. Discrepancies in peak height and shape are due to the signal intensities being given in transmittance for the exp spectra, and molar absorption coefficient ( ε ) for the theor spectra. To report the exp spectra with values of ε , amounts (i.e., film thickness or concentration) must be known and they are difficult to measure accurately using conventional spectral acquisition techniques. Likewise, to convert the theor spectra to absorbance, concentrations must be known, which are irrelevant for isolated molecules in the gas phase.
The small peak at 3814 cm 1 of the exp spectrum corresponds with the peak at 3642 cm 1 of the theor spectrum. In this case, mapping the exp signal to the theor signal is straightforward because the theor peak at 3642 cm 1 is the highest normal vibrational mode, and the only signal close to 3814 cm 1 , with the nearest theor signal at 3086 cm 1 . Mapping exp to theor frequencies of lower resolution is less straightforward, but can be done by comparing IR intensities, using common functional group frequency characteristics, and comparing distances between signals. In the 3058 to 2510 cm 1 domain of the exp spectrum, there are six weak signals. The theor spectrum shows only one peak at 3086 cm 1 , but there are three more weak signals from 3093 through 3062 cm 1 , which have low intensities at the given scale. These theor peaks correspond with the exp peaks in the domain of 3058 through 2809 cm 1 . In both spectra, there are 14 prominent signals between 1800 and 400 cm 1 .
Figure 4a,b show the exp and theor IR spectra for IPT. There is an intense, sharp signal at 3507 cm 1 of the theor spectrum, which is not pronounced in the exp spectrum. This is likely due to the effects of intermolecular hydrogen bonding (HB) in the solid phase. Between 2900 and 2300 cm 1 of the exp spectrum, there is a multitude of signals, whereas there is only a single peak at 3084 cm 1 in the theor spectrum. As in the case of IBA for IBA, there are several low-intensity theor signals in the 3200 to 2800 cm 1 domain. On the other hand, none of the exp signals between 2700 and 2300 cm 1 are pronounced in the theor spectrum; this is also the case for the broad signal at 1823 cm 1 . From 1700 through 400 cm 1 , the signals match up well between exp and theor spectra.
Figure 5a,b are the exp and theor IR spectra of BIO-1, respectively. BIO-1 is a I-hydroxylated HV derivative of the monovalent IBA. The exp and theor spectra of IBA are quite similar to the spectra of BIO-1. However, the spectra of IBA and BIO-1 diverge in the 3100 to 2400 cm 1 domain, where BIO-1 has fewer signals and no pronounced signals in the 3600 to 4000 cm 1 domain. In addition, the prominent signals at 733 and 1250 cm 1 of IBA have higher relative frequencies for BIO-1. Notably, the latter signal is nearly 90 cm 1 higher in BIO-1. For the theor spectrum of BIO-1, the peaks at 3701 and 3083 cm 1 have higher frequencies by 59 and 3 cm 1 , respectively, compared to the corresponding signals of IBA. On the other hand, many of the signals between 1800 and 400 cm 1 have lower frequencies in BIO-1 compared to IBA.
Figure 6a,b show the exp and theor IR spectra of BIAT-5, respectively. The exp spectrum has six broad peaks between 3500 and 2400 cm 1 and several weak signals between 2400 and 1500 cm 1 . There are two signals at 3394 and 3463 cm 1 which are not pronounced in the spectrum of IPT. The other three peaks at 3062, 2970, and 2901 cm 1 are more than 100 cm 1 higher than corresponding signals for IPT. At 2408 cm 1 , there is a single peak, whereas IPT has several signals in the same domain. There is an absence of pronounced signals between 1900 and 1500 cm 1 , but this domain contains noise which may be significant. There are three weak signals from 1454 to 1377 cm 1 , which are almost 20 cm 1 lower than the corresponding signals of IPT. The theor spectrum of BIAT-5 contains sharp peaks at 3697 and 3085 cm 1 , with the former having high intensity. The theor and exp spectra of BIAT-5 match well, with the exception being the exp signal at 2408 cm 1 , which is not pronounced in the theor spectrum. Comparison between theor spectra of BIAT-5 and IPT reveals all of the signals between 3697 and 1235 cm 1 have higher frequencies in BIAT-5, with the peak at 3697 cm 1 being 190 cm 1 higher compared to the corresponding peak of IPT. Conversely, between 1235 and 400 cm 1 , IPT has higher frequencies, the exception being at 738 cm 1 in BIAT-5 and 736 cm 1 in IPT.
There are significant differences between exp and theor IR spectra which are mostly attributable to the phase difference in this case, but it is evident that the two can be mapped to one another. LVM analysis not only confirms the accurate mapping of exp to theor vibrational frequencies, but it also provides a probe of vibrational spectra with a level of detail otherwise unobtainable, as is demonstrated in the following subsection.

CNM Analysis

There are many common IR signatures and numerous charts exist to help with assigning functional groups for the most recognizable signals, but doing so neglects the fact that normal vibrational modes are delocalized. In other words, most vibrational frequencies are composed of several molecular motions rather than a single C−C or C−N stretch for example. This is where LVM theory provides valuable new insights for the analysis of vibrational spectra through decomposition of normal vibrational modes into individual LVM contributions from local parameters bond distance (R), angle ( θ ), and dihedral/torsion ( φ ).
Figure 7 shows CNM, for IBA, including normal mode frequencies 439 through 3507 cm 1 . IBA has eight normal modes with frequencies below 400 cm 1 , but the CNM plots do not include these frequencies in accordance with the exp IR spectrum. Most of these eight frequencies are related to the LVMs of iodine, and this is consistent for CNM of IPT, BIO-1, and BIAT-5. In addition, included in Figure 7 is a schematic showing the atom numbers and a legend for the color coding of individual LVM contributions: R, θ , and φ .
There are three highly localized normal modes: (1) 3642 cm 1 (3810 cm 1 exp) which is the pure O−H stretch, (2) 78%, 1752 cm 1 (1668 cm 1 exp) is the C12-O13 carbonyl stretch combined with bending motions of C3-C12/O−C−O, and (3) 576 cm 1 (635 cm 1 exp) is 93% dihedral C3-C12-O14-H15 wagging motion. The normal vibrational mode at 451 cm 1 corresponds with the 465 cm 1 peak in the exp spectrum, and is composed primarily of C3−C12−O13 and C3−C12−O14 bending modes (56%), which consist of the Ar C-atom bound to COOH, both O-atoms, and the COOH C-atom. Vibrations from the bending motions of C3−C4−I11, C3−C12, and C2−C3−C12 account for 17% of this normal mode. C3 (the Ar C-atom) is the only atom included in each of the aforementioned LVMs. Other notable signals are at 733/738 (theor/exp), 1013/1002, and 1250/1170 cm 1 . In addition, 44% of the signal at 1170/1250 cm 1 is comprised of the C12-O14-H15 bend—10% and 9% of this mode consist of the C3-C12 stretch and the C12-O14 asymmetric stretch. The signals at 3058, 2972, and 2809 cm 1 (3086 cm 1  theor) are the C−H symmetric stretches.
The IR spectra of 5-phenyltetrazole derivatives with various substituents in the phenyl rings, including the spectrum of the ortho-iodo derivative, IPT, have been reported [138]. Six of the absorption bands in the spectrum of IPT, listed as 1467(s), 1269(w), 1048(s), 1018(m), 1005(s), and 767(m) cm 1 , were assigned to vibrations of the TZ ring. Each of these bands were also observed in this work (shown in italics in Figure 4a), and are within 3 to 4 cm 1 of the previous literature values. The authors in [138] provide an in-depth discussion on the IR spectrum of 5-phenyltetrazole and the combinations of vibrations responsible for the occurrence of all absorption bands are in agreement with the above assignments. For example, the band at 1465 cm 1 in the spectrum of 5-phenyltetrazole (likely of similar origin as the band at 1467 (1471 in this work) cm 1 ) in the spectrum of IPT was attributed to a combination of vibrations from C−NH (37%), C=N (36%), N−NH (16%), and N=N (15%). In addition, more complex vibrations involving three atoms, such as N−N=N (12%) and C=N−N (9%), were reported [139]. In the current work, LVM contributions to the frequencies observed in the IR spectrum of IPT are discussed in more detail.
Figure 8 is the CNM plot of IPT, including normal modes at 439 through 3507 cm 1 . The highest theor frequency at 3507 cm 1 is a pure N−H stretch. This peak is not in the exp spectrum, which is likely due to intermolecular hydrogen bonding in the solid phase. No other normal modes for IPT contain more than 59% contributions from a single local mode.
Figure 9 is the CNM plot of BIO-1, including frequencies spanning 418 through 3701 cm 1 . The highest theor frequency at 3701 cm 1 is a pure N−H stretch; this peak is not pronounced in the exp spectrum, which is again attributable to intermolecular HBs in the solid phase. Important LVM contributions are as follows: 3081/3059 (exp) → 3093 (theor), symmetric ortho-C10−H13 and m e t a -C9−H15; 3083 (theor), symmetric stretch m e t a -C7−H14, p a r a -C8−H12. The other two C−H stretches are asymmetric; 3073, asymmetric stretch of the two m e t a -C−H bonds C7−H14 ( m e t a ), C9−H15 ( m e t a ), C8−H12 ( p a r a ), plus ortho-C10-H13; 3061: asymmetric stretch of the two m e t a -C−H bonds m e t a -C7−H14, C9−H15 ( m e t a ), C8−H12 ( p a r a ).
Figure 10 is the CNM plot of BIAT-5, including normal modes spanning 426 through 3697 cm 1 . The theor signal at 3697 cm 1 , which corresponds with exp frequencies of 3463 and 3394 cm 1 , is a pure O−H stretch. Theor frequencies 3093, 3085, and 3075 cm 1 (3062, 2970, 2901 cm 1 exp) consist entirely of C−H symmetric and asymmetric stretches. The exp signal at 2408 cm 1 is not pronounced in the theor spectrum. The exp signal at 1454 cm 1 corresponds with the theor signal at 1447 cm 1 . Motions combined within this mode are: C−H wag (11% m e t a -C10−H16, 10% C12−C10−H16) and C−N asymmetric stretches (17% N3−C13, N4−C13). The remaining 67% of this frequency consists of numerous Ar C−C and C−H stretches, N−C−N angular modes, and C9−C13 bend/C11−C9−C13 angular modes. Each of the aforementioned LVMs contributes < 8 % . The theor signal at 1381 cm 1 (1377 cm 1 exp) corresponds with N3−C13 stretch (26%) and N4−C13 bend (7%), 8% N2−N3−C13, and 30% ortho-C8−H15/ m e t a -C10−H16/ p a r a -C12−H18 bends including C−C−H. 1117 (1128 exp) cm 1 consists primarily of N−N asymmetric stretches (47%) and N−N−N angular mode (46%). The signal at 920 cm 1 (830 exp) is a pure I1−O6−H14 bend. The normal mode at 738 cm 1 (737 exp) is 14% I1−O6−H14 and 76% O6−I1−C7−C8, and at 514 (579 exp) cm 1 is 93% I1−O6 and 426 (431 exp) is 94% O6−I1−C7−C8 torsion. The normal modes at 426, 514, 517, and 695 through 920 cm 1 mainly consist of the various LVMs associated with iodine. The N−N and N−N−N LVMs are mostly contained within normal modes 995, 1080, and 1117 cm 1 .

3.2. NBO Analysis

In Figure 1, atomic and group NBO charges for all molecules (reference systems IBA and IPT, BIOs 14, and BIATs 58) investigated in this work are reported (see also Table 3). NBO atomic charges of iodine are shown in purple, carbon in black, oxygen in red, and nitrogen in blue. Figure 1 also shows the total NBO charge of the Ar group in black, the TZ group in blue, and the alkoxy group in red. The iodine-atomic NBO charge in IBA is slightly more positive compared to IPT (0.179 and 0.158 e, respectively) and the total charge of the Ar group in IBA is significantly more negative compared to IPT (−0.213 and −0.126 e, respectively); while the total charges of COOH in IBA and TZ in IPT are both close to zero, with COOH having a net positive charge (0.035 e) and TZ having a net negative charge (−0.032 e).
Comparing NBO charges of IBA with those of IPT reveals some significant electronic differences which extend to BIOs 14 and BIATs 58, namely the following: charge (and electronic density) distribution are more polar in IBA compared to IPT, and NBO charge/electronic density is significantly higher/more negative on the Ar ring in IBA compared to IPT. On the other hand, in IPT, more electron density is transferred to TZ, which becomes more negative relative to the total COOH charge in IBA, showing a stronger electron withdrawing effect of the TZ group. It is also interesting to note that the presence of the TZ group in IPT increases the electron density/negative charge on the Ar-carbon atom bound to iodine relative to that of IBA (NBO charges of −0.133 and −0.129 e, respectively). Consequently, iodine is less positive in IPT compared to IBA (NBO charge values of 0.158 and 0.179 e, respectively).
To reiterate some of the aforementioned observations, 18 show similar trends in charge and electron density distribution with respect to the comparison of IBA and IPT. The iodine-atomic charge in BIO-1 is slightly more positive compared to BIAT-5 (1.329 and 1.294 e, respectively), the charge of the Ar group is significantly more negative (−0.224 e BIO and −0.120 e BIAT, respectively), and the charge on TZ in 5 is more negative than COO in 1 (−0.639 and −0.569 e, respectively). However, in contrast with IBA and IPT, the Ar-C atomic charge (bound to iodine) in 1 is slightly more negative relative to 5 (−0.195 and −0.178 e, respectively). COO groups in BIOs 24 show similar charge differences compared with TZ groups in BIATs 6, 7, and 8, which show a similar increase in electron withdrawing abilities of the TZ groups in the BIATs, as was observed in IPT.
The electron withdrawing ability of the TZ groups in BIATs 58 increases in series, as the charges of these groups become more negative (−0.639, −0.642, −0.644, and −0.648 e, respectively), which correlates to more negative charges on the Ar groups (−0.120, −0.125, −0.128, and −0.130 e, respectively). A methyl group is a stronger σ -electron donor than H, and the increasing number of methyl substituents in the alkoxy groups MeO → EtO → iPrO increases electron density transfer to the Ar and TZ groups in series. This causes increased negative charges, despite the fact that both the TZ and the Ar groups are not directly bound to the alkoxy groups. It is also interesting to note that, in the same series, the charge on iodine becomes more positive (1.294, 1.302, 1.307, and 1.308 e, respectively). A similar trend is observed for BIOs 14. Increasing the number of carbon atoms in the alkoxy groups results in increasingly more negative charges on the COO groups (−0.569, −0.571, −0.573, and −0.575 e, respectively), and the Ar groups (−0.224, −0.228, −0.231, and −0.232 e, respectively), while the iodine-atomic charges become more positive (1.329, 1.336, 1.341, and 1.343 e, respectively). The iodine atoms in this case, similar to I2, show both σ -electron donor and acceptor capabilities [140]. In addition to the special properties of HVI, iodine can form an efficient bridge between molecular groups of different electron donor/acceptor abilities.
NBO charges of the hydroxyl/alkoxy O-atoms vary significantly in series for BIOs 14 (−1.003, −0.786, −0.795, −0.802 e, respectively), whereas the combined charge of the whole functional group only varies by 0.002 e in the same series. Likewise, for BIATs 58, O-atomic charges vary by 0.216 e, whereas hydroxyl/alkoxy group charges vary by 0.004 e. The I-atomic charge becomes more positive in the same series, i.e., the I-atomic charge increases in magnitude as the mass of the hydroxyl/alkoxy group increases. NBO charge of the Ar rings become slightly more negative in the same series (−0.224 to −0.232 e in BIOs 14 and −0.120 to −0.130 e in BIATs 58, respectively). With the exception of the hydroxyl/alkoxy O-atoms, the difference in atomic NBO charges between O, N, and C-atoms bound to iodine remains relatively constant. The NBO charge on iodine varies, but to a lesser extent compared to he hydroxyl/alkoxy O-atoms; Δ NBO = 0.014  e and 0.218 e, respectively. Generally, more insights into the differences in electronic structure, density, and populations are obtained by considering the net NBO charge of the entire functional group rather than the charge on the atom which bonds directly to iodine.
Most notable is the difference in NBO charges of the whole Ar rings between BIOs and BIATs. For the BIOs, Ar NBO charges are more negative compared to BIATs by up to 0.104 e (60.5% difference). On the other hand, BIAT TZ group charges are more negative compared to BIO OCO group charges by up to 0.079 e. This is a result of the ability of TZs, particularly pronounced in amino-TZs to withdraw electrons [141,142,143]. Iodine can also accept an electronic charge from an Ar ring through a σ -hole interaction. However, given the abundance of lone pairs electrons (lp) in iodine and comparatively low electronegativity, iodine is as likely to be an electron donor as it is an acceptor. This result seems subtle, but it points to the C−C bond between TZ/OCO and Ar as a channel for the transfer of negative charge from the latter to the former. This finding naturally draws interest in the LVMs as well as electronic density ( ρ ( r b ) ) of this connecting C−C bond. The importance of this connector C−C bond is addressed in the following subsection.

3.3. LVM and QTAIM Analyses

Table 3 depicts chemical bond strength of all iodine-containing bonds: I−C, I−N, and I−O, expressed as LVM stretching force constants k a , along with ionic character described by the local energy density H ( r b ) . In Figure 11a,b, k a of all I−N, I−O, and I−C bonds are plotted with respect to bond length R (Figure 11a) and H ( r b ) (Figure 11b). Generally, increase in bond strength roughly correlates with shorter R and increase in covalent character of the bond ( H ( r b ) values increasing in magnitude). The I−C and I−OR bonds are significantly stronger than the I−N and the I−O (CO) bonds. The I−C bonds are approximately 0.1 Å longer than the I−OR bonds, but are similar in strength. All eight I−OR bonds are with in 0.06 mDyn/Å of each other. The ten I−C bonds show the following trend in terms of bond strength: BIOs 14 > IBA > IPT > BIATs 58. The I−N bonds are slightly weaker than the I−O (CO) bonds, and bond strength decreases in series (i.e., bond strength decreases as size of OR increases).
In Figure 12a,b, k a of all I−N, I−O, and I−C bonds are plotted with respect to electronic density at the bond critical point (BCP) ρ ( r b ) (Figure 12a) and angle θ of the L−I−L 3c-4e bonds (Figure 12b). Similar to the trend of H ( r b ) , electron density increases with increasing bond strength. The I−N bonds have the lowest electron densities at the BCPs, and the I−C bonds of BIOs 14 have the most electron density at the BCPs. Increase in angle θ of the 3c-4e bond roughly correlates with increasing bond strength. This correlation is strongest for the I−N and I−O (CO) bonds. Interestingly, even though the I−C bonds of BIOs 14 are perpendicular to the 3c-4e bonds, increasing the O−I−O angle roughly translates to increase in I−C bond strength. This indicates that a more linear 3c-4e bond allows for better orbital overlap and more electronic density at the equatorial positions around iodine. In addition, it is notable that the 3c-4e bond angles in the BIATs are approximately 2.5 smaller compared to the BIOs.
Figure 13a,b show local vibrational force constants of the iodine-related bonds plotted against the difference of the NBO charges between the two bonding atoms. There is a linear correlation between Δ NBO charge and force constant for the I−N (TZ) and the I−O (OCO) bonds, with the stronger bonds having smaller Δ NBO charges. Compared to the BIOs, the BIATs have weaker iodine-related bonds and larger Δ NBO values. For the I−C bonds, there is a similar trend, but it is extremely subtle. There is no such correlation for the I−O (hydroxyl/alkoxy) bonds.
For BIOs 14, strength of the I−C bond decreases in series ( k a = 2.615, 2.590, 2.586, and 2.550 mDyn/Å, respectively), the energy density decreases almost in series (the H ( r b ) value of −0.544, −0.542, −0.542, and −0.540 Hartree/Å 3 , respectively), and the charge difference increases in series ( Δ NBO values of 1.524, 1.531, 1.536, and 1.538 e, respectively). In other words, the I−C bond in this series becomes weaker, less covalent, with increased polarization of the electron density. Strength of the I−O (hydroxyl/alkoxy) bonds do not follow the trend of the I−C bonds, with k a values of 2.572, 2.442, 2.462, and 2.459 mDyn/Å, respectively (I−OH > I−OEt > I−OiPr > I−OMe), H ( r b ) values of −0.425, −0.442, −0.446, and −0.444 Hartree/Å 3 , respectively (I−OEt > I−OiPr > I−OMe > I−OH), and Δ NBO values of 2.332, 2.122, 2.136, and 2.138 e, respectively (I−OH > I−OiPr > I−OEt > I−OMe). To summarize some rather surprising results for the I−O (hydroxyl/alkoxy) bonds, the I−OH bond is the strongest but has the lowest energy density, the I−OiPr bond is weaker than the I−OEt bond, and the I−OMe bond is the weakest of the four bonds. Excluding the I−N bonds of BIATs 58, the I−OCO bonds between iodine and the carboxylate ester group in BIOs 14 are the weakest of the iodine-related bonds ( k a values of 1.671, 1.609, 1.575, and 1.569 mDyn/Å, respectively), the energy density decreases in magnitude in series ( H ( r b ) values of −0.276, −0.273, −0.270, and −0.266 Hartree/Å 3 , respectively), and the charge difference increases ( Δ NBO values 2.069, 2.076, 2.082, and 2.085 e, respectively). These changes indicate that the I−OCO bond becomes weaker with less covalent character and larger more polarized electron density.
For BIATs 58, the I−C bond strength varies unpredictably ( k a values of 2.374, 2.356, 2.371, and 2.370 mDyn/Å, respectively), as does energy density ( H ( r b ) values of −0.504, −0.502, −0.503, and −0.501 Hartree/Å 3 , respectively), but the charge difference increases in series ( Δ NBO values of 1.472, 1.480, 1.485, and 1.486 e, respectively). This indicates that the polar character of the I−C bond increases in the same series. In BIAT-5, I−OH bond strength is slightly stronger than the I−OMe bond in BIAT-6 ( + 0.003 mDyn/Å), but generally bond strength increases in series ( k a values of 2.384, 2.381, 2.437, and 2.442 mDyn/Å, respectively). Energy density is 0.001 Hartree/Å 3 larger for BIAT-7 compared to BIAT-8, but generally increases in magnitude in series ( H ( r b ) values of −0.425, −0.443, −0.447, and −0.446 Hartree/Å 3 , respectively), and the I−OH bond of BIAT-5 is the most polar but charge difference increases in series for the three I−O (alkoxy) bonds ( Δ NBO values of 2.298, 2.090, 2.104, and 2.112 e, respectively), showing that the I−O (alkoxy) bond becomes stronger, more covalent, and less polarized. Bond strength of I−N in BIATs 58 decreases in series ( k a values of 1.431, 1.364, 1.325, and 1.307 mDyn/Å, respectively), the energy density decreases in series ( H ( r b ) values of −0.246, −0.241, −0.237, and −0.232 Hartree/Å 3 , respectively), and the charge difference increases in series ( Δ NBO values of 1.718, 1.725, 1.730, and 1.731 e, respectively),
Table 4 summarizes LVM, QTAIM, and NBO results for the C−C bonds which link Ar and OCO/TZ groups in IBA, IPT, BIOs 14, and BIATs 58. In addition to the I−N bonds, this bridging C−C bond is a potential weak point, as well as a location for electron transfer. These C−C bonds are significantly stronger than the iodine-related bonds but are significantly weaker than the C−C, C−O, C−N, and N−N bonds. According to Table 4, the C−C bond is stronger for IPT compared to IBA (the k a values of 4.468 and 4.408 mDyn/Å, respectively), is shorter (R values of 1.469 and 1.495 Å, respectively), and has a more covalent character (the H ( r b ) values of −1.782 and −1.688 Hartree/Å 3 ). The difference between the NBO charges of both C atoms in this bond for IPT is smaller compared to IBA ( Δ NBO values of 0.4368 and 0.9691 e, respectively), which indicates less polarization of this bond in IPT and is consistent with larger covalent character [144]. The chemical differences of the C−C bonds in IPT and IBA is explained by more π -electron delocalization in the TZ group of IPT, which leads to a less positive charge of the TZ-carbon compared to the OCO-carbon atom of IBA (NBO charge values are 0.3025 and 0.7700 e, respectively), resulting in the C−C bond in IPT being less polar but stronger. Similar correlations are observed in calculations for BIOs 14 and BIATs 58. The C−C bonds in BIATs 58 are generally stronger, shorter, and have more covalent character (correlating with a less polar bond) compared to BIOs 14, and the TZ-carbon atoms have less positive NBO charges in contrast with the OCO-carbon atoms. It is also interesting to note that the C−C ’bridge’ bond in BIAT-5 is much stronger than that of IPT, where both systems have TZ groups ( k a values of 5.587 and 4.468 mDyn/Å, respectively), thus indicating more π -electron delocalization in 5. This is corroborated by the TZ-carbon atom charge in BIAT-5 less positive compared with the IPT-carbon atom (NBO charges of 0.2483 and 0.3025 e, respectively), and the bridging C−C bond being less polar in BIAT-5 compared with IPT ( Δ NBO values of 0.3509 and 0.4368 e, respectively). The analogous C−C bonds in BIATs 68 are very similar to BIAT-5 in terms of all of the aforementioned properties, indicating that formation of the I−N bond in the BIATs increases the π electron delocalization in the TZ group, making these C−C bonds stronger.

4. Conclusions and Outlook

There are fundamental differences between exp and theor IR spectroscopy which are mostly attributable to the phase difference in this case, but it is evident that the two can be mapped to one another with relative ease if corrections to the harmonic approximation are considered. LVM analysis not only confirms the accurate mapping of exp to theor vibrational frequencies, but opens the door to the analysis of vibrational spectra at a level of detail and accuracy not previously known. This work demonstrates the importance of the often overlooked fact that normal vibrational modes are delocalized. LVM theory takes this into consideration, and reveals valuable new insights by decomposing normal vibrational modes into fundamental LVM contributions in terms of R, θ , and φ .
The sensitivity of LVM stretching force constants has also revealed subtle but important differences in the intrinsic nature of iodine-related bonds between a set of four alkoxybenziodoxoles, which are relatively stable, and four tetrazolated derivatives which decompose explosively. The I−N bonds in BIATs 58 are generally weaker than the comparable I−OCO bonds in BIOs 14 by as much as 0.300 mDyn/Å. Moreover, strength of the I−N bond in BIATs 58 decreases with the increasing size of the hydroxyl/alkoxy group: OH > MeO > EtO > iPrO. Based on LVM analysis, I−N is most likely the trigger bond responsible for initiation of explosive decomposition in BIATs 58. For comparison, trigger bonds were recently identified as C−N bonds in the high-energy compounds 1-fluoro-2,4,6-trinitrobenzene and 2,4,6-trinitrobenzaldehyde [31]. Our calculations predict the C−C bond linking Ar and aldehyde groups to be the weakest in the latter, but both bonds in question have k a values between 3.400 and 3.900 mDyn/Å, whereas the possible I−N trigger bonds, I−C, and I−O bonds in this work have k a values between 1.300 and 2.620 mDyn/Å. With regard to the I−N bonds, it is interesting to note that explosive compounds with I−N bonds have been known as long as elemental iodine itself. The formation of the highly friction-sensitive black compound NI3 (or more precisely, its adduct with ammonia), achieved by mixing iodine with aqueous ammonia, was first reported by Bernard Courtois in 1813 in the paper where the discovery of the new element was announced [145]. The composition of the unstable compound was confirmed as NI3·NH3 in 1903 [146]. Based on strength of the I−C, I−N, and I−O bonds, where stronger bonds will correlate with increase in stability, we predict stability of BIATs 58 as follows: BIAT-8 < BIAT-7 < BIAT-6 < BIAT-5. The finer details obtained from this work regarding the characteristics of iodine-containing bonds can lead to new structure–property relationships. Some additional key differences between the BIOs 14 and BIATs 58 are as follows:
  • The TZ groups in BIATs 58 are more capable electron withdrawing groups compared to the OCO groups in BIOs 14
  • Compared to BIATs 58, BIOs 14 have Ar rings nearly two times more negative in charge, iodine atoms with positive in charges, and stronger I−C and I−O (hydroxyl/alkoxy) bonds
  • The weakest bonds in this work are the I−N bonds
  • The 3c-4e L−I−L bond angles of BIATs 58 are up to 2.5 larger compared to the analogous bonds in BIOs 14
Future goals are to investigate the effects other types of substituents will have on the physical properties of benziodazolotetrazoles, which encompasses modifications to the TZ group as well as the Ar group. Furthermore, we plan to investigate the mechanism of detonation using the Unified Reaction Valley Approach to reveal in what order the most important changes in electronic structure occur and determine the most prevalent reaction products. Lastly, we wish to investigate and develop better methods for approximating the anharmonic nature of vibrational frequencies.

Supplementary Materials

The following are available at https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/physchem1010004/s1, Single Crystal X-ray Crystallography data of I-hydroxybenziodoxolone (BIO-1) including refinement details; Figure S1: Displacement ellipsoid plot (50% probability level) of (BIO-1) at 100.0(3) K; Table S1: Crystal data and structure refinement; Table S2: Fractional Atomic Coordinates and Equivalent Isotropic Displacement Parameters; Table S3: Anisotropic Displacement Parameters; Table S4: Bond length; Table S5: Bond Angles; Table S6: Torsion Angles; Table S7: Hydrogen Atom Coordinates.

Author Contributions

Conceptualization, E.K., N.V.T., and S.Y.; methodology, E.K., K.D.S., N.V.T., M.F., and S.Y.; validation, S.Y., M.F., and K.D.S.; formal analysis, S.Y., M.F., and K.D.S.; resources, E.K. and N.V.T.; writing—original draft preparation, S.Y. and E.K.; writing—review and editing, E.K., S.Y., M.F., K.D.S., and N.V.T.; visualization, S.Y., M.F., and K.D.S.; supervision, E.K. and N.V.T.; funding acquisition, E.K. and N.V.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Science Foundation Grant Nos. CHE-1464906 (E.K.) and CHE-1455200 (N.V.T.).

Acknowledgments

The authors thank SMU for generously providing computational resources and analytical instrumentation and Daniel Unruh of Texas Tech University for collection and analysis of the X-ray diffraction data. This work was completed while N.V.T. was serving at the National Science Foundation, and the support of the Foundation is gratefully acknowledged.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
3c-4e3-Center 4-Electron
ArOrtho-Disubstituted Benzene
BIATBenziodazolotetrazole
BIOAlkoxybenziodoxole
CNMCharacterization of Normal Modes
CTCharge Transfer
DFTDensity Functional Theory
expExperimental
HVHypervalent
HVIHypervalent Iodine
IBA2-Iodobenzoic Acid
IPT5-(2-Iodophenyl)tetrazole
LLigand
LVMLocal Vibrational Mode
MAEMean Absolute Error
NBONatural Bond Orbital
OCOCarboxyl/Carboxylate Ester
RCN4Tetrazolate
RCO2Carboxylate
RMSERoot Mean Square Error
theorTheoretical
TZTetrazole

References

  1. Varvoglis, A. Organic Compounds of Polycoordinated Iodine; Wiley-VCH: Weinheim, Germany, 1992; p. 414. ISBN 0471188255. [Google Scholar]
  2. Willgerodt, C. Die Organischen Verbindungen mit Mehrwertigem Jod; Ferdinand Enke: Stuttgart, Germany, 1914; p. 266. ISBN 978-0526136322. [Google Scholar]
  3. Varvoglis, A. Hypervalent Iodine in Organic Synthesis; Academic Press: San Diego, CA, USA, 1997; p. 223. ISBN 0-12-714975-9. [Google Scholar]
  4. Wirth, T. Hypervalent Iodine Chemistry: Modern Developments in Organic Synthesis; Springer: Berlin, Germany, 2003; p. 264. ISBN 978-3-540-44107-6. [Google Scholar]
  5. Zhdankin, V.V. Hypervalent Iodine Chemistry: Preparation, Structure and Synthetic Applications of Polyvalent Iodine Compounds; Wiley: Chichester, UK, 2014; p. 468. ISBN 9781118341032. [Google Scholar]
  6. Yoshimura, A.; Zhdankin, V.V. Advances in Synthetic Applications of Hypervalent Iodine Compounds. Chem. Rev. 2016, 116, 3328–3435. [Google Scholar] [CrossRef] [PubMed]
  7. Zhdankin, V.V. Hypervalent iodine compounds: Reagents of the future. Arkivoc 2020, iv, 1–11. [Google Scholar] [CrossRef] [Green Version]
  8. Zhdankin, V.V. Benziodoxole-Based Hypervalent Iodine Reagents in Organic Synthesis. Curr. Org. Synth. 2005, 2, 121–145. [Google Scholar] [CrossRef]
  9. Bamberger, J.; Ostler, F.; Mancheno, O.G. Frontiers in Halogen and Chalcogen-Bond Donor Organocatalysis. ChemCatChem 2019, 11, 5198–5211. [Google Scholar] [CrossRef]
  10. Vaish, A.; Tsarevsky, N.V. Hypervalent Iodine Compounds in Polymer Science and Technology. In Main Group Strategies towards Functional Organic Materials; Baumgartner, T., Jaekle, F., Eds.; Wiley: Chichester, UK, 2018; pp. 483–514. [Google Scholar] [CrossRef]
  11. Musher, J.I. The Chemistry of Hypervalent Molecules. Angew. Chem. Int. Ed. 1969, 8, 54–68. [Google Scholar] [CrossRef]
  12. Pimentel, G.C. The Bonding of Trihalide and Bifluoride Ions by the Molecular Orbital Method. J. Chem. Phys. 1951, 19, 446–448. [Google Scholar] [CrossRef]
  13. Hach, R.J.; Rundle, R.E. The Structure of Tetramethylammonium Pentaiodide. J. Am. Chem. Soc. 1951, 73, 4321–4324. [Google Scholar] [CrossRef]
  14. Munzarova, M.L.; Hoffmann, R. Electron-Rich Three-Center Bonding: Role of s,p Interactions across the p-Block. J. Am. Chem. Soc. 2002, 124, 4787–4795. [Google Scholar] [CrossRef]
  15. Durrant, M.C. A quantitative definition of hypervalency. Chem. Sci. 2015, 6, 6614–6623. [Google Scholar] [CrossRef] [Green Version]
  16. Kalemosa, A.; Ariyarathnab, I.R.; Khanb, S.N.; Miliordosb, E.; Mavridis, A. Hypervalency and the chemical bond. Comput. Theoret. Chem. 2019, 1153, 65–74. [Google Scholar] [CrossRef]
  17. Crabtree, R.H. Hypervalency, secondary bonding and hydrogen bonding: Siblings under the skin. Chem. Soc. Rev. 2017, 46, 1720–1729. [Google Scholar] [CrossRef]
  18. Kupwade, R.V. A Concise Review of Hypervalent Iodine with Special Reference to Dess- Martin Periodinane. Mini-Rev. Org. Chem. 2020, 17, 946-1–946-12. [Google Scholar] [CrossRef]
  19. Yannacone, S.; Oliveira, V.; Verma, N.; Kraka, E. A Continuum from Halogen Bonds to Covalent Bonds: Where Do λ3 Iodanes Fit? Inorganics 2019, 7, 47. [Google Scholar] [CrossRef] [Green Version]
  20. Varvoglis, A. Aryliodine(III) Dicarboxylates. Chem. Soc. Rev. 1981, 10, 377. [Google Scholar] [CrossRef]
  21. Waser, J. Benziodoxol(on)e Reagents as Tools in Organic Synthesis: The Background behind the Discovery at the Laboratory of Catalysis and Organic Synthesis. Synlett 2016, 27, 2761–2773. [Google Scholar] [CrossRef] [Green Version]
  22. Ostrovskii, V.A.; Koldobskii, G.I.; Trifonov, R.E. Tetrazoles. In Comprehensive Heterocyclic Chemistry III; Katritzky, A.R., Ramsden, C.A., Scriven, E.F.V., Taylor, R.J.K., Eds.; Elsevier: Oxford, UK, 2008; pp. 257–423. [Google Scholar] [CrossRef]
  23. Kumar, R.; Vaish, A.; Runčevski, T.; Tsarevsky, N.V. Hypervalent Iodine Compounds with Tetrazole Ligands. J. Org. Chem. 2018, 83, 12496–12506. [Google Scholar] [CrossRef] [PubMed]
  24. Vaish, A.; Sayala, K.D.; Tsarevsky, N.V. N-Heterocycle (Tetrazole)-Stabilized Pseudocyclic λ3-Iodane: Synthesis and Reactivity. Tetrahedron Lett. 2019, 60, 150995. [Google Scholar] [CrossRef]
  25. Kumar, R.; Sayala, K.D.; Camdzic, L.L.; Siegler, M.; Vaish, A.; Tsarevsky, N. Heterocyclic Hypervalent Iodine(III) Compounds with Fused Benziodazole and Tetrazole Rings (I-Substituted Tetrazolo[1,5-b][1,2]Benziodazoles). ChemRxiv 2020. [Google Scholar] [CrossRef]
  26. Wang, T.; Gao, H.; Shreeve, J.M. Functionalized Tetrazole Energetics: A Route to Enhanced Performance. Z. Anorg. Allg. Chem. 2021, 647, 157–191. [Google Scholar] [CrossRef]
  27. Singh, R.P.; Verma, R.D.; Meshri, D.T.; Shreeve, J.M. Energetic Nitrogen-Rich Salts and Ionic Liquids. Angew. Chem. Int. Ed. 2006, 45, 3584–3601. [Google Scholar] [CrossRef] [PubMed]
  28. Trifonov, R.E.; Ostrovskii, V.A. Protolytic Equilibria in Tetrazoles. Russ. J. Org. Chem. 2006, 42, 1585–1605. [Google Scholar] [CrossRef]
  29. Wani, M.Y.; Silva, M.R.; Krishnakumar, B.; Kumar, S.; Al-Bogami, A.S.; Aqlan, F.M.; Sobral, A.J.F.N. Catalytic Synthesis of 5-Substituted Tetrazoles: Unexpected Reactions and Products. J. Heterocycl. Chem. 2019, 56, 1613–1621. [Google Scholar] [CrossRef]
  30. Koch, E.C. High Explosives, Propellants, Pyrotechnics; DeGruyter: Boston, MA, USA, 2021. [Google Scholar]
  31. Dosch, D.E.; Reichel, M.; Born, M.; Klapötke, T.M.; Karaghiosoff, K. Investigation of Structure-Property Relationships of Three Nitroaromatic Compounds: 1-Fluoro-2,4,6-Trinitrobenzene, 2,4,6-Trinitrophenyl Methanesulfonate, and 2,4,6-Trinitrobenzaldehyde. Cryst. Growth Des. 2020, 21, 243–248. [Google Scholar] [CrossRef]
  32. Li, J. Relationships for the Impact Sensitivities of Energetic C-Nitro Compounds Based on Bond Dissociation Energy. J. Phys. Chem. B 2010, 114, 2198–2202. [Google Scholar] [CrossRef] [PubMed]
  33. Li, J. A Quantitative Relationship for the Shock Sensitivities of Energetic Compounds based on X−NO2 (X = C, N, O) Bond Dissociation Energy. J. Hazard. Mater. 2010, 180, 768–772. [Google Scholar] [CrossRef] [PubMed]
  34. Luo, Y.R. Comprehensive Handbook of Chemical Bond Energies; Taylor and Francis: Boca Raton, FL, USA, 2007. [Google Scholar]
  35. Morse, M.D. Predissociation measurements of bond dissociation energies. Acc. Chem. Res. 2018, 52, 119–126. [Google Scholar] [CrossRef]
  36. Zhang, J.; Zhang, Q.; Vo, T.T.; Parrish, D.A.; Shreeve, J.M. Energetic Salts with π-Stacking and Hydrogen-Bonding Interactions Lead the Way to Future Energetic Materials. J. Am. Chem. Soc. 2015, 137, 1697–1704. [Google Scholar] [CrossRef] [PubMed]
  37. Cremer, D.; Kraka, E. From Molecular Vibrations to Bonding, Chemical Reactions, and Reaction Mechanism. Curr. Org. Chem. 2010, 14, 1524–1560. [Google Scholar] [CrossRef]
  38. Kalescky, R.; Kraka, E.; Cremer, D. Are Carbon-Halogen Double and Triple Bonds Possible? Int. J. Quantum Chem. 2014, 114, 1060–1072. [Google Scholar] [CrossRef]
  39. Kalescky, R.; Zou, W.; Kraka, E.; Cremer, D. Quantitative Assessment of the Multiplicity of Carbon-Halogen Bonds: Carbenium and Halonium Ions with F, Cl, Br, and I. J. Phys. Chem. A 2014, 118, 1948–1963. [Google Scholar] [CrossRef]
  40. Oliveira, V.; Kraka, E.; Cremer, D. Quantitative Assessment of Halogen Bonding Utilizing Vibrational Spectroscopy. Inorg. Chem. 2016, 56, 488–502. [Google Scholar] [CrossRef] [PubMed]
  41. Setiawan, D.; Sethio, D.; Cremer, D.; Kraka, E. From Strong to Weak NF Bonds: On the Design of a New Class of Fluorinating Agents. Phys. Chem. Chem. Phys. 2018, 20, 23913–23927. [Google Scholar] [CrossRef] [PubMed]
  42. Sethio, D.; Oliveira, V.; Kraka, E. Quantitative Assessment of Tetrel Bonding Utilizing Vibrational Spectroscopy. Molecules 2018, 23, 2763. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Grant, D.J.; Matus, M.H.; Switzer, J.R.; Dixon, D.A.; Francisco, J.S.; Christe, K.O. Bond Dissociation Energies in Second-Row Compounds. J. Phys. Chem. A 1966, 112, 3145–3156. [Google Scholar] [CrossRef] [PubMed]
  44. Konkoli, Z.; Cremer, D. A New Way of Analyzing Vibrational Spectra. I. Derivation of Adiabatic Internal Modes. Int. J. Quantum Chem. 1998, 67, 1–9. [Google Scholar] [CrossRef]
  45. Konkoli, Z.; Larsson, J.A.; Cremer, D. A New Way of Analyzing Vibrational Spectra. II. Comparison of Internal Mode Frequencies. Int. J. Quantum Chem. 1998, 67, 11–27. [Google Scholar] [CrossRef]
  46. Konkoli, Z.; Cremer, D. A New Way of Analyzing Vibrational Spectra. III. Characterization of Normal Vibrational Modes in terms of Internal Vibrational Modes. Int. J. Quantum Chem. 1998, 67, 29–40. [Google Scholar] [CrossRef]
  47. Konkoli, Z.; Larsson, J.A.; Cremer, D. A New Way of Analyzing Vibrational Spectra. IV. Application and Testing of Adiabatic Modes within the Concept of the Characterization of Normal Modes. Int. J. Quantum Chem. 1998, 67, 41–55. [Google Scholar] [CrossRef]
  48. Cremer, D.; Larsson, J.A.; Kraka, E. New Developments in the Analysis of Vibrational Spectra on the Use of Adiabatic Internal Vibrational Modes. In Theoretical and Computational Chemistry; Parkanyi, C., Ed.; Elsevier: Amsterdam, The Netherlands, 1998; pp. 259–327. [Google Scholar]
  49. Kraka, E.; Zou, W.; Tao, Y. Decoding Chemical Information from Vibrational Spectroscopy Data: Local Vibrational Mode Theory. WIREs Comput. Mol. Sci. 2020, 10, e1480. [Google Scholar] [CrossRef]
  50. Klapötke, T.M. Chemistry of High-Energy Materials, 5th ed.; DeGruyter: Boston, MA, USA, 2019. [Google Scholar]
  51. Kettner, M.A.; Klapötke, T.M. New Energetic Polynitrotetrazoles. Chem. Eur. J. 2015, 21, 3755–3765. [Google Scholar] [CrossRef]
  52. Köhler, J.; Meyer, R.; Homburg, A. Explosivstoffe; Wiley: Weinheim, Germany, 2008. [Google Scholar]
  53. Becke, A.D. Density-Functional Thermochemistry. III. The Role of Exact Exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef] [Green Version]
  54. Becke, A.D. A New Mixing of Hartree-Fock and Local Density-Functional Theories. J. Chem. Phys. 1993, 98, 1372–1377. [Google Scholar] [CrossRef]
  55. Becke, A.D. Density-Functional Thermochemistry. V. Systematic Optimization of Exchange-Correlation Functionals. J. Chem. Phys. 1997, 107, 8554–8560. [Google Scholar] [CrossRef]
  56. Salomon, O.; Reiher, M.; Hess, B.A. Assertion and Validation of the Performance of the B3LYP Functional for the First Transition Metal Row and the G2 Test Set. J. Chem. Phys. 2002, 117, 4729–4737. [Google Scholar] [CrossRef]
  57. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A Consistent and Accurate Ab Initio Parametrization of Density Functional Dispersion Correction (DFT-D) for the 94 Elements H-Pu. J. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the Damping Function in Dispersion Corrected Density Functional Theory. J. Comput. Chem. 2011, 32, 1456–1465. [Google Scholar] [CrossRef]
  59. Zhao, Y.; Truhlar, D.G. The M06 Suite of Density Functionals for Main Group Thermochemstry, Thermochemical Kinetics, Noncovalent Interactions, Excited States, and Transition Elements: Two New Functionals and Systematic Testing of Four M06-Class Functionals and 12 Other Functionals. Theor. Chem. Acc. 2007, 120, 215–241. [Google Scholar]
  60. Chai, J.D.; Head-Gordon, M. Long-Range Corrected Hybrid Density Functionals with Damped Atom-Atom Dispersion Corrections. Phys. Chem. Chem. Phys. 2008, 10, 6615–6620. [Google Scholar] [CrossRef] [Green Version]
  61. Dunning, T.H. Gaussian Basis Sets for use in Correlated Molecular Calculations. I. The Atoms Boron through Neon and Hydrogen. J. Chem. Phys. 1989, 90, 1007–1023. [Google Scholar] [CrossRef]
  62. Kendall, R.A.; Dunning, T.H.; Harrison, R.J. Electron Affinities of the First-Row Atoms Revisited. Systematic Basis Sets and Wave Functions. J. Chem. Phys. 1992, 96, 6796–6806. [Google Scholar] [CrossRef] [Green Version]
  63. Woon, D.E.; Dunning, T.H. Gaussian Basis Sets for use in Correlated Molecular Calculations. III. The Atoms Aluminum through Argon. J. Chem. Phys. 1993, 98, 1358–1371. [Google Scholar] [CrossRef] [Green Version]
  64. Woon, D.E.; Dunning, T.H. Gaussian Basis Sets for use in Correlated Molecular Calculations. IV. Calculation of Static Electrical Response Properties. J. Chem. Phys. 1994, 100, 2975–2988. [Google Scholar] [CrossRef] [Green Version]
  65. Peterson, K.A. Systematically Convergent Basis Sets with Relativistic Pseudopotentials. I. Correlation Consistent Basis Sets for the Post-d Group 13-15 Elements. J. Chem. Phys. 2003, 119, 11099–11112. [Google Scholar] [CrossRef] [Green Version]
  66. Peterson, K.A.; Figgen, D.; Goll, E.; Stoll, H.; Dolg, M. Systematically Convergent Basis Sets with Relativistic Pseudopotentials. II. Small-Core Pseudopotentials and Correlation Consistent Basis Sets for the Post-d Group 16-18 Elements. J. Chem. Phys. 2003, 119, 11113–11123. [Google Scholar] [CrossRef] [Green Version]
  67. Schäfer, A.; Horn, H.; Ahlrichs, R. Fully optimized contracted Gaussian basis sets for atoms Li to Kr. J. Chem. Phys. 1992, 97, 2571–2577. [Google Scholar] [CrossRef]
  68. Schäfer, A.; Huber, C.; Ahlrichs, R. Fully optimized contracted Gaussian basis sets of triple zeta valence quality for atoms Li to Kr. J. Chem. Phys. 1994, 100, 5829–5835. [Google Scholar] [CrossRef]
  69. Weigend, F.; Ahlrichs, R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297. [Google Scholar] [CrossRef]
  70. Weigend, F. Accurate Coulomb-fitting basis sets for H to Rn. Phys. Chem. Chem. Phys. 2006, 8, 1057–1065. [Google Scholar] [CrossRef]
  71. Heinen, F.; Engelage, E.; Cramer, C.J.; Huber, S.M. Hypervalent Iodine(III) Compounds as Biaxial Halogen Bond Donors. J. Am. Chem. Soc. 2020, 142, 8633–8640. [Google Scholar] [CrossRef]
  72. Sibert III, E.L. Modeling vibrational anharmonicity in infrared spectra of high frequency vibrations of polyatomic molecules. J. Chem. Phys. 2019, 150, 090901. [Google Scholar] [CrossRef]
  73. Morse, M.D. Clusters of Transition-Metal Atoms. Chem. Rev. 1986, 86, 1049–1109. [Google Scholar] [CrossRef]
  74. Byrd, E.F.C.; Sherrill, C.D.; Head-Gordon, M. The Theoretical Prediction of Molecular Radical Species: A Systematic Study of Equilibrium Geometries and Harmonic Vibrational Frequencies. J. Phys. Chem. A 2001, 105, 9736–9747. [Google Scholar] [CrossRef]
  75. Coolidge, M.B.; Marlin, J.E.; Stewart, J.J.P. Calculations of Molecular Vibrational Frequencies using Semiempirical Methods. J. Comput. Chem. 1991, 12, 948–952. [Google Scholar] [CrossRef]
  76. Galabov, B.; Yamaguchi, Y.; Remington, R.B.; Schaefer, H.F. High Level Ab Initio Quantum Mechanical Predictions of Infrared Intensities. J. Phys. Chem. A 2002, 106, 819–832. [Google Scholar] [CrossRef]
  77. Halls, M.D.; Velkovski, J.; Schlegel, H.B. Harmonic Frequency Scaling Factors for Hartree-Fock, S-VWN, B-LYP, B3-LYP, B3-PW91 and MP2 with the Sadlej pVTZ Electric Property Basis Set. Theor. Chem. Acc. 2001, 105, 413–421. [Google Scholar] [CrossRef]
  78. Irikura, K.K.; Johnson, R.D.; Kacker, R.N. Uncertainties in Scaling Factors for Ab Initio Vibrational Frequencies. J. Phys. Chem. A 2005, 109, 8430–8437. [Google Scholar] [CrossRef]
  79. Scott, A.P.; Radom, L. Harmonic Vibrational Frequencies: An Evaluation of Hartree-Fock, Møller-Plesset, Quadratic Configuration Interaction, Density Functional Theory, and Semiempirical Scale Factors. J. Phys. Chem. 1996, 100, 16502–16513. [Google Scholar] [CrossRef]
  80. Shefter, E.; Wolf, W. Crystal and Molecular Structure of 1,3Dihydro-1-hydroxy-3-oxo-1,2-benziodoxole. Nature 1964, 203, 512–513. [Google Scholar] [CrossRef]
  81. Shefter, E.; Wolf, W. Crystal and Molecular Structure of 1,3Dihydro-1-hydroxy-3-oxo-1,2-benziodoxole. J. Pharm. Sci. 1965, 54, 104–107. [Google Scholar] [CrossRef]
  82. Gräfenstein, J.; Cremer, D. Efficient DFT integrations by locally augmented radial grids. J. Chem. Phys. 2007, 127, 164113. [Google Scholar] [CrossRef] [Green Version]
  83. Huang, Y.; Chang, C.; Yuan, J.; Zhao, Z. High-Harmonic and Terahertz Spectroscopy (HATS): Methods and Applications. Appl. Sci. 2019, 9, 853. [Google Scholar] [CrossRef] [Green Version]
  84. Wu, G. Vibrational Spectroscopy; De Gruyter: Berlin, Germany, 2019. [Google Scholar]
  85. Smith, E.; Dent, G. Modern Raman Spectroscopy: A Practical Approach; Wiley: New York, NY, USA, 2019. [Google Scholar]
  86. Wilson, E.B.; Decius, J.C.; Cross, P.C. Molecular Vibrations; McGraw-Hill: New York, NY, USA, 1955. [Google Scholar]
  87. Wilson, E.B., Jr. A Method of Obtaining the Expanded Secular Equation for the Vibration Frequencies of a Molecule. J. Chem. Phys. 1939, 7, 1047–1052. [Google Scholar] [CrossRef]
  88. Woodward, L.A. Introduction to the Theory of Molecular Vibrations and Vibrational Spectroscopy; Oxford University Press: Oxford, UK, 1972. [Google Scholar]
  89. Quack, M. Spectra and Dynamics of Coupled Vibrations in Polyatomic Molecules. Annu. Rev. Phys. Chem. 1990, 41, 839–874. [Google Scholar] [CrossRef]
  90. Halonen, L. Local Mode Vibrations in Polyatomic Molecules. In Advances in Chemical Physics; Prigogine, I., Rice, S., Eds.; Wiley: New York, NY, USA, 1998; Volume 104, pp. 41–179. [Google Scholar]
  91. Jensen, P. An introduction to the theory of local mode vibrations. Mol. Phys. 2000, 98, 1253–1285. [Google Scholar] [CrossRef]
  92. Zou, W.; Cremer, D. C2 in a Box: Determining its Intrinsic Bond Strength for the X1Σ+g Ground State. Chem. Eur. J. 2016, 22, 4087–4097. [Google Scholar] [CrossRef]
  93. Oomens, J.; Kraka, E.; Nguyen, M.K.; Morton, T.M. Structure, Vibrational Spectra, and Unimolecular Dissociation of Gaseous 1-Fluoro-1-phenethyl Cations. J. Phys. Chem. A 2008, 112, 10774–10783. [Google Scholar] [CrossRef] [PubMed]
  94. Zou, W.; Kalescky, R.; Kraka, E.; Cremer, D. Relating Normal Vibrational Modes to Local Vibrational Modes: Benzene and Naphthalene. J. Mol. Model. 2012, 19, 2865–2877. [Google Scholar] [CrossRef] [PubMed]
  95. Kalescky, R.; Kraka, E.; Cremer, D. Identification of the Strongest Bonds in Chemistry. J. Phys. Chem. A 2013, 117, 8981–8995. [Google Scholar] [CrossRef]
  96. Kalescky, R.; Kraka, E.; Cremer, D. Description of Aromaticity with the Help of Vibrational Spectroscopy: Anthracene and Phenanthrene. J. Phys. Chem. A 2013, 118, 223–237. [Google Scholar] [CrossRef] [PubMed]
  97. Kalescky, R.; Kraka, E.; Cremer, D. New Approach to Tolman’s Electronic Parameter Based on Local Vibrational Modes. Inorg. Chem. 2013, 53, 478–495. [Google Scholar] [CrossRef]
  98. Humason, A.; Zou, W.; Cremer, D. 11,11Dimethyl-1,6-methano[10]annulene-An Annulene with an Ultralong CC Bond or a Fluxional Molecule? J. Phys. Chem A. 2014, 119, 1666–1682. [Google Scholar] [CrossRef] [PubMed]
  99. Sethio, D.; Lawson Daku, L.M.; Hagemann, H.; Kraka, E. Quantitative Assessment of B−B−B, B−Hb−B, and B−Ht Bonds: From BH3 to B12H122−. ChemPhysChem 2019, 20, 1967–1977. [Google Scholar] [CrossRef]
  100. Makoś, M.Z.; Freindorf, M.; Sethio, D.; Kraka, E. New Insights into Fe-H2 and Fe-H Bonding of a [NiFe] Hydrogenase Mimic–A Local Vibrational Mode Study. Theor. Chem. Acc. 2019, 138, 76. [Google Scholar] [CrossRef]
  101. Makoś, M.Z.; Zou, W.; Freindorf, M.; Kraka, E. Metal-Ring Interactions in Actinide Sandwich Compounds: A Combined Normalized Elimination of the Small Component and Local Vibrational Mode Study. Mol. Phys. 2020, 118, e1768314. [Google Scholar] [CrossRef]
  102. Verma, N.; Tao, Y.; Zou, W.; Chen, X.; Chen, X.; Freindorf, M.; Kraka, E. A Critical Evaluation of Vibrational Stark Effect (VSE) Probes with the Local Vibrational Mode Theory. Sensors 2020, 20, 2358. [Google Scholar] [CrossRef] [Green Version]
  103. Freindorf, M.; Kraka, E. Critical Assessment of the FeC and CO Bond strength in Carboxymyoglobin–A QM/MM Local Vibrational Mode Study. J. Mol. Model. 2020, 26, 281-1–281-15. [Google Scholar] [CrossRef] [PubMed]
  104. Kraka, E.; Freindorf, M. Characterizing the Metal Ligand Bond Strength via Vibrational Spectroscopy: The Metal Ligand Electronic Parameter (MLEP). In Topics in Organometallic Chemistry–New Directions in the Modeling of Organometallic Reactions; Lledós, A., Ujaque, G., Eds.; Springer: Berlin/Heidelberg, Germany, 2020; Volume 67, pp. 1–43. [Google Scholar]
  105. Freindorf, M.; Kraka, E.; Cremer, D. A Comprehensive Analysis of Hydrogen Bond Interactions Based on Local Vibrational Modes. Int. J. Quantum Chem. 2012, 112, 3174–3187. [Google Scholar] [CrossRef]
  106. Kalescky, R.; Zou, W.; Kraka, E.; Cremer, D. Local Vibrational Modes of the Water Dimer–Comparison of Theory and Experiment. Chem. Phys. Lett. 2012, 554, 243–247. [Google Scholar] [CrossRef]
  107. Kalescky, R.; Kraka, E.; Cremer, D. Local Vibrational Modes of the Formic Acid Dimer–The Strength of the Double H-Bond. Mol. Phys. 2013, 111, 1497–1510. [Google Scholar] [CrossRef]
  108. Kraka, E.; Freindorf, M.; Cremer, D. Chiral Discrimination by Vibrational Spectroscopy Utilizing Local Modes. Chirality 2013, 25, 185–196. [Google Scholar] [CrossRef]
  109. Setiawan, D.; Kraka, E.; Cremer, D. Description of Pnicogen Bonding with the help of Vibrational Spectroscopy-The Missing Link between Theory and Experiment. Chem. Phys. Lett. 2014, 614, 136–142. [Google Scholar] [CrossRef]
  110. Setiawan, D.; Kraka, E.; Cremer, D. Strength of the Pnicogen Bond in Complexes Involving Group VA Elements N, P, and As. J. Phys. Chem. A 2014, 119, 1642–1656. [Google Scholar] [CrossRef]
  111. Setiawan, D.; Kraka, E.; Cremer, D. Hidden Bond Anomalies: The Peculiar Case of the Fluorinated Amine Chalcogenides. J. Phys. Chem. A 2015, 119, 9541–9556. [Google Scholar] [CrossRef] [PubMed]
  112. Kraka, E.; Setiawan, D.; Cremer, D. Re-Evaluation of the Bond Length-Bond Strength Rule: The Stronger Bond Is not Always the Shorter Bond. J. Comp. Chem. 2015, 37, 130–142. [Google Scholar] [CrossRef] [PubMed]
  113. Zhang, X.; Dai, H.; Yan, H.; Zou, W.; Cremer, D. B-H π Interaction: A New Type of Nonclassical Hydrogen Bonding. J. Am. Chem. Soc. 2016, 138, 4334–4337. [Google Scholar] [CrossRef]
  114. Setiawan, D.; Cremer, D. Super-Pnicogen Bonding in the Radical Anion of the Fluorophosphine Dimer. Chem. Phys. Lett. 2016, 662, 182–187. [Google Scholar] [CrossRef] [Green Version]
  115. Oliveira, V.; Kraka, E.; Cremer, D. The Intrinsic Strength of the Halogen Bond: Electrostatic and Covalent Contributions Described by Coupled Cluster Theory. Phys. Chem. Chem. Phys. 2016, 18, 33031–33046. [Google Scholar] [CrossRef]
  116. Tao, Y.; Zou, W.; Jia, J.; Li, W.; Cremer, D. Different Ways of Hydrogen Bonding in Water–Why Does Warm Water Freeze Faster than Cold Water? J. Chem. Theory Comput. 2017, 13, 55–76. [Google Scholar] [CrossRef]
  117. Oliveira, V.; Cremer, D. Transition from Metal-Ligand Bonding to Halogen Bonding Involving a Metal as Halogen Acceptor: A Study of Cu, Ag, Au, Pt, and Hg Complexes. Chem. Phys. Lett. 2017, 681, 56–63. [Google Scholar] [CrossRef]
  118. Oliveira, V.; Cremer, D.; Kraka, E. The Many Facets of Chalcogen Bonding: Described by Vibrational Spectroscopy. J. Phys. Chem. A 2017, 121, 6845–6862. [Google Scholar] [CrossRef] [PubMed]
  119. Oliveira, V.; Kraka, E. Systematic Coupled Cluster Study of Noncovalent Interactions Involving Halogens, Chalcogens, and Pnicogens. J. Phys. Chem. A 2017, 121, 9544–9556. [Google Scholar] [CrossRef] [PubMed]
  120. Zou, W.; Zhang, X.; Dai, H.; Yan, H.; Cremer, D.; Kraka, E. Description of an Unusual Hydrogen Bond Between Carborane and a Phenyl Group. J. Organometal. Chem. 2018, 856, 114–127. [Google Scholar] [CrossRef]
  121. Lyu, S.; Beiranvand, N.; Freindorf, M.; Kraka, E. Interplay of Ring Puckering and Hydrogen Bonding in Deoxyribonucleosides. J. Phys. Chem. A 2019, 123, 7087–7103. [Google Scholar] [CrossRef]
  122. Oliveira, V.P.; Marcial, B.L.; Machado, F.B.C.; Kraka, E. Metal-Halogen Bonding Seen through the Eyes of Vibrational Spectroscopy. Materials 2020, 13, 55. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Tao, Y.; Qiu, Y.; Zou, W.; Nanayakkara, S.; Yannacone, S.; Kraka, E. In Situ Assessment of Intrinsic Strength of X-I⋯OA Type Halogen Bonds in Molecular Crystals with Periodic Local Vibrational Mode Theory. Molecules 2020, 25, 1589. [Google Scholar] [CrossRef] [Green Version]
  124. Yannacone, S.; Sethio, D.; Kraka, E. Quantitative Assessment of Intramolecular Hydrogen Bonds in Neutral Histidine. Theor. Chem. Acc. 2020, 139, 125-1–125-10. [Google Scholar] [CrossRef]
  125. Martins, J.; Quintino, R.P.; Politi, J.R.S.; Sethio, D.; Gargano, R.; Kraka, E. Computational Analysis of Vibrational Frequencies and Rovibrational Spectroscopic Constants of Hydrogen Sulfide Dimer using MP2 and CCSD(T). Spectrochim. Acta A 2020, 239, 118540-1–118540-9. [Google Scholar] [CrossRef]
  126. Yannacone, S.; Freindorf, M.; Tao, Y.; Zou, W.; Kraka, E. Local Vibrational Mode Analysis of π-Hole Interactions between Aryl Donors and Small Molecule Acceptors. Crystals 2020, 10, 556. [Google Scholar] [CrossRef]
  127. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian16 Revision B.01; Gaussian Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  128. Zou, W.; Tao, Y.; Freindorf, M.; Makoś, M.Z.; Verma, N.; Kraka, E. Local Vibrational Mode Analysis (LModeA); Computational and Theoretical Chemistry Group (CATCO), Southern Methodist University: Dallas, TX, USA, 2020. [Google Scholar]
  129. Reed, A.E.; Curtiss, L.A.; Weinhold, F. Intermolecular Interactions from a Natural Bond Orbital, Donor-Acceptor Viewpoint. Chem. Rev. 1988, 88, 899–926. [Google Scholar] [CrossRef]
  130. Weinhold, F.; Landis, C.R. Valency and Bonding: A Natural Bond Orbital Donor-Acceptor Perspective; Cambridge University Press: Cambridge, UK, 2005. [Google Scholar]
  131. Weinhold, F.; Landis, C.R.; Glendening, E.D. What is NBO Analysis and How is it Useful? Int. Rev. Phys. Chem. 2016, 35, 39–440. [Google Scholar] [CrossRef]
  132. Keith, T.A. AIMAll (Version 17.11.14). 2017. Available online: aim.tkgristmill.com (accessed on 12 May 2021).
  133. Cremer, D.; Kraka, E. Chemical Bonds without Bonding Electron Density? Does the Difference Electron-Density Analysis Suffice for a Description of the Chemical Bond? Angew. Chem. Int. Ed. 1984, 23, 627–628. [Google Scholar] [CrossRef]
  134. Cremer, D.; Kraka, E. A Description of the Chemical Bond in Terms of Local Properties of Electron Density and Energy. Croat. Chem. Acta 1984, 57, 1259–1281. [Google Scholar]
  135. Kraka, E.; Cremer, D. Chemical Implication of Local Features of the Electron Density Distribution. In Theoretical Models of Chemical Bonding. The Concept of the Chemical Bond; Maksic, Z.B., Ed.; Springer: Berlin/Heidelberg, Germany, 1990; Volume 2, pp. 453–542. [Google Scholar]
  136. Kraszkiewicz, L.; Skulski, L. Optimized syntheses of iodylarenes from iodoarenes, with sodium periodate as the oxidant. Part II. Arkivoc 2003, 2003, 120–125. [Google Scholar] [CrossRef] [Green Version]
  137. Origin(Pro) 9.1.0 (64 bit); OriginLab Corporation: Northampton, MA, USA, 2013.
  138. Kaczmarek, J.; Grzonka, Z. Substituent Effects in Infrared Absorption Spectra of 5-Phenyltetrazoles. Pol. J. Chem. 1980, 54, 1297–1300. [Google Scholar]
  139. Matveeva, N.A.; Sushko, N.I.; Makarevich, N.I.; Gaponik, P.N.; Ivashkevich, O.A.; Koren, A.O. Substituent Effects in Infrared Absorption Spectra of 5-Phenyltetrazoles. Zh. Prikl. Spektrosk. 1992, 57, 442–450. [Google Scholar]
  140. Rogachev, A.Y.; Hoffmann, R. Iodine (I2) as a Janus-Faced Ligand in Organometallics. J. Am. Chem. Soc. 2013, 135, 3262–3275. [Google Scholar] [CrossRef] [PubMed]
  141. Boyer, J.H.; Miller, E.J. The Tetrazole-Azidoazomethine Equilibrium. J. Am. Chem. Soc. 1959, 81, 4671–4673. [Google Scholar] [CrossRef]
  142. Henry, R.A.; Finnegan, W.G.; Lieber, E. Kinetics of the Isomerization of Substituted 5-Aminotetrazoles. J. Am. Chem. Soc. 1955, 77, 2264–2270. [Google Scholar] [CrossRef]
  143. Lieber, E.; Rao, C.N.R.; Chao, T.S.; Ramachandran, J. Relative Acidities of 5-(Substituted Phenyl)amino-4-phenyl-1,2,3-triazoles. J. Org. Chem. 1958, 23, 1916–1918. [Google Scholar] [CrossRef]
  144. Sanderson, R. Polar Covalence; Academic Press: New York, NY, USA, 1983. [Google Scholar]
  145. Courtois, B. Découverte d’une substance nouvelle dans le varech. Ann. Chim. 1813, 88, 304–310. [Google Scholar]
  146. Silberrad, O. IX.–The Constitution of Nitrogen Iodide. J. Chem. Soc. Trans. 1905, 87, 55–66. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Schematic of reference molecules IBA and IPT, BIOs 14 and BIATs 58, showing atomic NBO charges (units in [e]) of iodine (purple), oxygen (red), carbon/hydrogen (black), and nitrogen (blue); calculated at the ω B97X-D/Def2-TZVP level of theory. Summed NBO charges of hydroxyl/alkoxy, carboxylate ester (OCO), tetrazole (TZ), and ortho-substituted benzene (Ar) groups are shown below each respective molecule. Note: For BIOs 24 and BIATs 68, the charges shown in black font to the left of molecule numbers are the total charges of the Me, Et, and iPr groups, respectively.
Figure 1. Schematic of reference molecules IBA and IPT, BIOs 14 and BIATs 58, showing atomic NBO charges (units in [e]) of iodine (purple), oxygen (red), carbon/hydrogen (black), and nitrogen (blue); calculated at the ω B97X-D/Def2-TZVP level of theory. Summed NBO charges of hydroxyl/alkoxy, carboxylate ester (OCO), tetrazole (TZ), and ortho-substituted benzene (Ar) groups are shown below each respective molecule. Note: For BIOs 24 and BIATs 68, the charges shown in black font to the left of molecule numbers are the total charges of the Me, Et, and iPr groups, respectively.
Physchem 01 00004 g001
Figure 2. Displacement ellipsoid plot (50% probability level) of I-hydroxybenziodoxolone (BIO-1) at 100.0(3) K (left) and intermolecular secondary bonding in the crystals of BIO-1 (center). Selected bond lengths, bond angles, and torsion angles are provided in the right hand-side panel.
Figure 2. Displacement ellipsoid plot (50% probability level) of I-hydroxybenziodoxolone (BIO-1) at 100.0(3) K (left) and intermolecular secondary bonding in the crystals of BIO-1 (center). Selected bond lengths, bond angles, and torsion angles are provided in the right hand-side panel.
Physchem 01 00004 g002
Figure 3. (a) Experimental, and (b) theoretical IR spectra of IBA.
Figure 3. (a) Experimental, and (b) theoretical IR spectra of IBA.
Physchem 01 00004 g003
Figure 4. (a) Experimental, and (b) theoretical IR spectra of IPT.
Figure 4. (a) Experimental, and (b) theoretical IR spectra of IPT.
Physchem 01 00004 g004
Figure 5. (a) Experimental and (b) theoretical IR spectra of BIO-1.
Figure 5. (a) Experimental and (b) theoretical IR spectra of BIO-1.
Physchem 01 00004 g005
Figure 6. (a) Experimental and (b) theoretical IR spectra of BIAT-5.
Figure 6. (a) Experimental and (b) theoretical IR spectra of BIAT-5.
Physchem 01 00004 g006
Figure 7. Decomposition of normal vibrational modes into % LVM contributions, also called characterization of normal modes (CNM) for IBA; normal vibrational modes 451 through 3642 cm 1 . Normal modes below 400 cm 1 are excluded to coordinate with experimental IR spectra.
Figure 7. Decomposition of normal vibrational modes into % LVM contributions, also called characterization of normal modes (CNM) for IBA; normal vibrational modes 451 through 3642 cm 1 . Normal modes below 400 cm 1 are excluded to coordinate with experimental IR spectra.
Physchem 01 00004 g007
Figure 8. CNM plot of IPT; normal vibrational modes in the frequency domain of 439 through 3507 cm 1 .
Figure 8. CNM plot of IPT; normal vibrational modes in the frequency domain of 439 through 3507 cm 1 .
Physchem 01 00004 g008
Figure 9. CNM plot of BIO-1; normal vibrational modes in the frequency domain of 418 through 3701 cm 1 .
Figure 9. CNM plot of BIO-1; normal vibrational modes in the frequency domain of 418 through 3701 cm 1 .
Physchem 01 00004 g009
Figure 10. CNM plot of BIAT-5; normal vibrational modes in the frequency domain of 426 through 3697 cm 1 .
Figure 10. CNM plot of BIAT-5; normal vibrational modes in the frequency domain of 426 through 3697 cm 1 .
Physchem 01 00004 g010
Figure 11. Local mode vibrational force constant k a plotted with respect to (a) R, and (b) H ( r b ) of iodine-based bonds for structures 18.
Figure 11. Local mode vibrational force constant k a plotted with respect to (a) R, and (b) H ( r b ) of iodine-based bonds for structures 18.
Physchem 01 00004 g011
Figure 12. Local vibrational mode stretching force constant k a of all iodine-based bonds (three per compound) plotted with respect to (a) ρ ( r b ) , and (b) bond angle θ of O−I−O (14) and O−I−N (58) axial−I−axial 3c-4e interactions.
Figure 12. Local vibrational mode stretching force constant k a of all iodine-based bonds (three per compound) plotted with respect to (a) ρ ( r b ) , and (b) bond angle θ of O−I−O (14) and O−I−N (58) axial−I−axial 3c-4e interactions.
Physchem 01 00004 g012
Figure 13. Local mode vibrational force constant k a of iodine-related bonds plotted with respect to Δ NBO charge between the bonding atoms (a) I−C (Ar) and I−O hydroxyl/alkoxy (R), and (b) I−N and I−O (OCO).
Figure 13. Local mode vibrational force constant k a of iodine-related bonds plotted with respect to Δ NBO charge between the bonding atoms (a) I−C (Ar) and I−O hydroxyl/alkoxy (R), and (b) I−N and I−O (OCO).
Physchem 01 00004 g013
Table 1. Comparison of experimental and theoretical values of the bond length R (Å), the bond angle θ (Deg), and the dihedral angle φ (Deg) for BIO-1.
Table 1. Comparison of experimental and theoretical values of the bond length R (Å), the bond angle θ (Deg), and the dihedral angle φ (Deg) for BIO-1.
Parameterexp a ω B97X-DB3LYPM06-2X
acpVTZTZVPacpVTZTZVPacpVTZTZVP
R (I−C)2.0982.1002.0912.1252.1162.0952.087
R (I−OR)1.9752.0232.0092.0542.0402.0162.004
R (I−OCO)2.2742.1312.1212.1602.1502.1242.113
θ (O−I−O)167.5169.7169.8169.9170.0169.1169.3
θ (C−I−OR)90.991.191.191.591.590.590.4
θ (C−I−OCO)76.878.678.878.378.578.778.9
φ (O−C−O−I)179.6179.0178.8179.1179.0178.8178.7
φ (C−C−O−I)−1.9−1.2−1.4−1.0−1.2−1.3−1.5
a X-ray structure determined in this work. Details are provided in the Supplementary Materials.
Table 2. Comparison of experimental (exp) IR frequencies ( ω ( e x p ) ) with theoretical (theor) normal mode vibrational frequencies ( ω μ ) for BIO- 1 .
Table 2. Comparison of experimental (exp) IR frequencies ( ω ( e x p ) ) with theoretical (theor) normal mode vibrational frequencies ( ω μ ) for BIO- 1 .
Model Chemistry% Error (Average)MAEStandard DeviationRMSE
ω B97X-D/acpVTZ-PP1.8116.4421.1326.35
ω B97X-D/TZVP1.5012.9916.2520.48
B3LYP/acpVTZ-PP1.8917.3118.3824.91
B3LYP/TZVP2.1119.0617.6725.69
B3LYP/TZVPP2.1218.8219.0226.41
B3LYP-D3/acpVTZ-PP1.8616.9518.0924.46
B3LYP-D3/TZVP2.0718.5417.5725.24
B3LYP-D3/TZVPP2.0518.0019.1025.89
B3LYP-D3BJ/acpVTZ-PP1.7716.3817.7323.81
B3LYP-D3BJ/TZVP1.8016.9516.3723.28
B3LYP-D3BJ/TZVPP1.9717.7618.2625.15
M06-2X/acpVTZ-PP1.4513.6116.1320.80
M06-2X/TZVP2.1722.0417.2527.72
Averages are based on 20 of the 21 exp IR frequencies (see below). MAE are given in cm−1.
Table 3. Theoretical bond lengths R (Å), local mode vibrational force constants k a (mDyn/Å), local mode vibrational frequencies ω a (cm 1 ), electronic densities at the bond critical points ρ ( r b ) (e/Å 3 ), energy densities at the bond critical points H ( r b ) (Hartree/Å 3 ), and atomic NBO charges (e) for all I−C, I−N, and I−O bonds in IBA, IPT, and 18, calculated at the ω B97X-D/Def2-TZVP level of theory.
Table 3. Theoretical bond lengths R (Å), local mode vibrational force constants k a (mDyn/Å), local mode vibrational frequencies ω a (cm 1 ), electronic densities at the bond critical points ρ ( r b ) (e/Å 3 ), energy densities at the bond critical points H ( r b ) (Hartree/Å 3 ), and atomic NBO charges (e) for all I−C, I−N, and I−O bonds in IBA, IPT, and 18, calculated at the ω B97X-D/Def2-TZVP level of theory.
#BondR k a ω a ρ ( r b ) H ( r b ) AtomCharge
IBAI−C42.0972.536626.60.887−0.488C4−0.199
------I+0.179
IPTI−C42.1022.468618.10.881−0.482C4−0.134
------I+0.158
1I−O22.1211.671446.80.679−0.276O2−0.740
I−O42.0092.472543.40.862−0.425O4−1.003
I−C2.0922.615636.30.939−0.544C11−0.195
------I+1.329
2I−O22.1251.609438.40.675−0.273O2−0.740
I−O42.0022.442540.10.880−0.442O4−0.786
I−C2.0932.590633.20.937−0.542C11−0.195
------I+1.336
3I−O22.1281.575433.80.671−0.270O2−0.741
I−O41.9992.462542.40.885−0.446O4−0.795
I−C2.0932.586632.70.937−0.542C11−0.195
------I+1.341
4I−O22.1311.552430.60.666−0.266O2−0.742
I−O41.9982.459542.10.884−0.444O4−0.802
I−C2.0932.574631.30.936−0.540C11−0.194
------I+1.343
5I−N2.2041.389432.30.636−0.246N3−0.424
I−O2.0102.434539.20.862−0.425O6−1.004
I−C2.1172.378606.80.904−0.504C7−0.178
------I+1.294
6I−N2.2091.333428.50.630−0.241N3−0.423
I−O2.0022.410533.40.881−0.443O6−0.788
I−C2.1172.360603.90.902−0.502C7−0.178
------I+1.302
7I−N2.2131.325422.30.624−0.237N3−0.423
I−O1.9992.437539.70.887−0.447O6−0.797
I−C2.1172.371605.90.903−0.503C7−0.178
------I+1.307
8I−N2.2181.307419.50.617−0.232N3−0.423
I−O1.9982.442540.20.887−0.446O6−0.804
I−C2.1182.370605.70.901−0.501C7−0.178
------I+1.308
Table 4. Theoretical R (Å), k a (mDyn/Å), ω a (cm 1 ), ρ ( r b ) (e/Å 3 ), H ( r b ) (Hartree/Å 3 ), and NBO atomic charges (e) for the C−C bonds linking Ar groups and TZ/OCO groups in IBA, IPT, and BIOs/BIATS 18.
Table 4. Theoretical R (Å), k a (mDyn/Å), ω a (cm 1 ), ρ ( r b ) (e/Å 3 ), H ( r b ) (Hartree/Å 3 ), and NBO atomic charges (e) for the C−C bonds linking Ar groups and TZ/OCO groups in IBA, IPT, and BIOs/BIATS 18.
#R k a ω a ρ ( r b ) H ( r b ) AtomChargeAtomCharge Δ Charge
IBA1.49504.408626.61.8291−1.6880C12 (OCO)0.7700C3 (Ar)−0.19910.9691
IPT1.46904.468618.11.8679−1.7820C15 (TZ)0.3025C3 (Ar)−0.13430.4368
11.50334.1231080.01.7951−1.6272C5 (OCO)0.7478C6 (Ar)−0.19030.9380
21.50374.1201079.61.7935−1.6246C5 (OCO)0.7482C6 (Ar)−0.19020.9385
31.50424.1041077.41.7920−1.6222C5 (OCO)0.7486C6 (Ar)−0.18970.9382
41.50424.1091078.11.7920−1.6222C5 (OCO)0.7484C6 (Ar)−0.18910.9375
51.44895.5871257.11.9489−1.9120C13 (TZ)0.2483C9 (Ar)−0.10260.3509
61.44895.5791256.21.9487−1.9116C8 (TZ)0.2467C4 (Ar)−0.10230.3490
71.44885.5881257.31.9486−1.9114C13 (TZ)0.2459C9 (Ar)−0.10190.3477
81.44895.5831256.71.9494−1.9127C12 (TZ)0.2448C8 (Ar)−0.10120.3461
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yannacone, S.; Sayala, K.D.; Freindorf, M.; Tsarevsky, N.V.; Kraka, E. Vibrational Analysis of Benziodoxoles and Benziodazolotetrazoles. Physchem 2021, 1, 45-68. https://0-doi-org.brum.beds.ac.uk/10.3390/physchem1010004

AMA Style

Yannacone S, Sayala KD, Freindorf M, Tsarevsky NV, Kraka E. Vibrational Analysis of Benziodoxoles and Benziodazolotetrazoles. Physchem. 2021; 1(1):45-68. https://0-doi-org.brum.beds.ac.uk/10.3390/physchem1010004

Chicago/Turabian Style

Yannacone, Seth, Kapil Dev Sayala, Marek Freindorf, Nicolay V. Tsarevsky, and Elfi Kraka. 2021. "Vibrational Analysis of Benziodoxoles and Benziodazolotetrazoles" Physchem 1, no. 1: 45-68. https://0-doi-org.brum.beds.ac.uk/10.3390/physchem1010004

Article Metrics

Back to TopTop