Abstract
In the present research, we focus on the energetics and electronic aspects of enhanced reactivity in the regioselective bioorthogonal 1,3-dipolar cycloaddition reaction of various substituted cyclooctynes with methyl azide, applying quantum chemistry approaches. In this respect, we assessed the structural and energetic properties of regioisomeric products and their corresponded transition states and calculated the reaction electronic energy changes and energy barriers through the cycloaddition pathways. The obtained results revealed that the trifluoromethyl substitution and fluorination of cyclooctynes lead to improved reactivity, in conjunction with increased exothermicity and decreased activation energy values. On the other hand, quantum theory of atoms in molecules computations were performed on some key bond and ring critical points that demonstrated the stabilizing topological properties of electron density and its derivatives upon trifluoromethyl substitution and fluorination of propargylic carbon of cyclooctynes which can be regarded as the essential origin of enhanced reactivity.
1 Introduction
Cyclooctyne-based compounds have recently attracted much attention in the bioorthogonal chemistry field as potential bioorthogonal reagents [1,2,3]. In recent decades, bioorthogonal reactions have been broadly applied in the metabolism of living systems without mixing with the main biological processes. Furthermore, bioorthogonal reactions are widely used in selective and rapid labeling of biomolecules for live-cell imaging, with no disruption of biological processes [4,5,6,7,8,9,10].
In this route, the bioorthogonal cycloaddition reaction of strained alkynes with common 1,3 dipoles such as azides has been investigated from the experimental and computational viewpoints [11,12,13,14,15,16].
Copper (Cu)-catalyzed azide–alkyne cycloaddition reactions introduced by Sharpless et al. [17] and Meldal et al. [18] independently lead to the regioselective and high-yield production 1,2,3-triazoles, with wide applications in organic synthesis, biochemical and drug discovery, chemical biology, and materials chemistry. However, due to the toxicity of copper that confines the application of Cu-catalyzed click reactions in biological scopes [19,20], the strain-promoted cycloaddition between cyclooctynes and azides were considered as a proper alternative process.
In this direction, many endeavors were allocated to interpret these cycloadditions-enhanced reactivities and selectivities quantitatively. From the biological point of view, it was shown that copper-free click cycloaddition of strained cycloalkynes with azides are fast enough for in vivo bioorthogonal labeling of biomolecules at physiological temperatures without the toxic side effects of copper [6,21–24]. Therefore, the researchers have concentrated on synthesizing highly strained cyclic alkynes, such as bicycle nonyne, azadibenzocyclooctyne, 4-dibenzocyclooctynol, difluorobenzocyclooctyne, and thiacycloheptyne [25–28].
On the other hand, to quantify the improved orthogonal reactivity of cycloalkynes, several computational investigations were performed on a series of cycloaddition reactions between 1,3-dipoles and cyclic alkynes as dipolarophiles using the distortion/interaction (D/I) model [29,30]. In this line, the enhancement of reactivity in cycloalkynes was modeled compared to unstrained alkynes, and it was shown that cycloalkynes require less distortion barrier to gain the transition-state structure, leading to both reducing the distortion energy and improving interaction energy.
In order to provide an accurate understanding of the effects of strain and electronics in reactivity, a series of cyclooctynes, benzocyclooctadienynes, dibenzo, and aza-analogues have been assessed via comparison with the calculated activation barriers [31], obtained by relative rate constants and D/I analysis using density functional theory (DFT) [32] methods.
In recent years, we have computationally focused on the energetic and electronic origins of regioselectivity in palladium- and ruthenium-catalyzed coupling reactions and azide–alkyne cycloadditions along the mechanical pathway [33,34] via DFT and quantum theory of atoms in molecules (QTAIM) [32,35,36] computations. Furthermore, we have investigated the stereoselective behavior of metal-free click synthesis of 1H-tetrazole-5-yl-acrylonitriles [37] from the mechanical and electronic viewpoints. With this in mind, we have presented a review article on the combined computational and experimental features of uncatalyzed Huisgen azide–alkyne cycloaddition, leading to the formation of 1,4 and 1,5-disubstituted 1,2,3-triazoles mixture [38,39].
In this context, in this research, we have focused on the probing of electronic aspects in cycloaddition of methyl azide with cyclooctyne and a range of substituted cyclooctynes at propargylic position to identify the stereoelectronic origin of strain-promoted rate enhancement in this type of 1,3-dipolar cycloadditions. To this goal, we comparatively assessed the structure and energetics of transition state barriers for the aforementioned concerted 1,3-dipolar cycloadditions and then analyzed the topological properties of electron density as the quantum mechanical observable quantity via QTAIM approach.
More clearly speaking, we concentrated on the topological analysis of electron density functions and their first (gradient) and second derivatives (Hessian), which revealed the zones of charge concentration and charge depletion for the newly formed bond critical points (BCPs) and ring critical points (RCPs) along the cycloaddition reaction path. In this line, we derived QTAIM quantified chemical information on the nature of strains and interactions in selectivity and reactivity of these types of azide–cyclooctyne cycloadditions.
2 Computational details
M08-HX functional methodology [40] in conjunction with 6-311+G** basis sets have been used to optimize the geometry of all reactants, products, and their corresponded transition states, all in the C1 symmetry point group. Furthermore, harmonic frequency calculations were performed for all stationary points to demonstrate the local minima and saddle points. All DFT computations have been performed using GAMESS suite of programs [41].
On the other hand, the obtained M08-HX/6-311+G** wave functions for all optimized structures were used as input by AIM2000 [42] and were analyzed by QTAIM approach. In this route, electron density (ρ
b), its Laplacian (
3 Results and discussion
3.1 Enhanced reactivity of fluorinated cyclooctynes in azide–cyclooctyne cycloaddition reaction: energetic aspect
As we discussed earlier, the main objective of this research is to investigate the stereoelectronic role of propargylic substitution of cyclooctyne by methyl, trifluoromethyl, and fluoride on the enhancement of reactivity in strain-promoted azide–cyclooctyne cycloaddition reactions (as illustrated in Figure 1)

Schematic representation of cycloaddition reaction of various substituted cyclooctynes with methyl azide, leading to produce two regioisomeric 1,2,3 triazoles.
In this respect, we primarily obtained the optimized geometry of all reactants, regioisomeric 1,4 and 1,5-disubstituted triazole products (denoted as isomer1 and isomer2, respectively) and their corresponding regioisomeric transition states (denoted as TS1 and TS2, respectively), followed by determination of activation energy barriers and the reaction electronic energies. In Figures 2–6, we have displayed reaction paths of cyclooctyne, methyl-substituted cyclooctyne, trifluoromethyl-substituted cyclooctyne, and mono and difluorinated cyclooctyne with methyl azide, which lead to producing regioselective 1,4 and 1,5-disubstituted 1,2,3-triazoles together with M08-HX/6-311+G** optimized geometry of all species.

Cycloaddition reaction pathway of methyl azide with cyclooctyne, in association with the optimized structure of triazole product and the corresponded transition state. We have reported the M08-HX/6-311+G** calculated values of activation energies and reaction electronic energies in the figure (note that the units are in kcal/mol).

Cycloaddition reaction pathway of methyl azide with propargylic substituted cyclooctyne by methyl, in association with the optimized structure of regioisomeric triazole products and the corresponded transition states. We have reported the M08-HX/6-311+G** calculated values of activation energies and reaction electronic energies in the figure (note that the units are in kcal/mol).

Cycloaddition reaction pathway of methyl azide with propargylic two-substituted cyclooctyne by trifluoromethyl, in association with the optimized structure of regioisomeric triazole products and the corresponded transition states. We have reported the M08-HX/6-311+G** calculated values of activation energies and reaction electronic energies in the figure (note that the units are in kcal/mol).

Cycloaddition reaction pathway of methyl azide with propargylic mono-substituted cyclooctyne by fluoride, in association with the optimized structure of regioisomeric triazole products and the corresponded transition states. We have reported the M08-HX/6-311+G** calculated values of activation energies and reaction electronic energies in the figure (note that the units are in kcal/mol).

Cycloaddition reaction pathway of methyl azide with propargylic two-substituted cyclooctyne by fluoride, in association with the optimized structure of regioisomeric triazole products and the corresponded transition states. We have reported the M08-HX/6-311+G** calculated values of activation energies and reaction electronic energies in the figure (note that the units are in kcal/mol).
The M08-HX/6-311+G** calculated values of activation energies and the reaction electronic energies for 1,3-dipolar cycloaddition of cyclooctyne, methyl-substituted cyclooctyne, and mono and di-fluorinated cyclooctynes with methyl azide are reported in Figures 2–6.
The following three facts can be extracted from the comparative analysis of the reported result in Figures 2–6: (1) In all mentioned 1,3-dipolar cycloaddition reactions, the synthesis of both regioisomers is exothermic with the small difference between the calculated reaction electronic energy values (about 1–2 kcal/mol) and reveals the requisiteness for considering the energetics of transition states in two regioselective mechanistic reaction pathways. (2) Transition state barrier for the production of isomer2 is about 3 kcal/mol lower than the synthesis of isomer1, which may be rationally regarded as the energetic symptom for the regioselective behavior. (3) Fluorination adjacent to the triple bond of cyclooctyne increases the exothermicity and decreases the activation energy values which is in agreement with the reported rate enhancement computations and elucidations [13,14].
On the other hand, methyl and trifluoromethyl substitution at propargylic carbon of cyclooctyne show decrease in exothermicity, about 3 and 6 kcal/mol, respectively. While methyl substitution led to a rise in the barrier energy values, in an opposite trend, trifluoromethyl substitution reduced the activation energy barriers. Clearly speaking, changing from non-substituted cyclooctyne to trifluoromethyl-substituted and monofluorinated cyclooctynes decreases the transition state barrier by about 2 kcal/mol, and in the case of difluorinated cyclooctyne, this substitution effect on reducing energy barriers is more intense by about 5 kcal/mol. Meanwhile, methyl substitution has no significant effect on the rate enhancement of these cycloaddition reactions. It should be noted that the regioselective behavior of the synthesis has been preserved via substituting cyclooctyne at a tertiary carbon α position.
3.2 QTAIM interpretations for the enhanced reactivity
In this section, we concentrated on QTAIM topological analysis of electron density and its derivatives [38,39] via the quantum mechanical structures, which is the collection of BCPs and RCPs, respectively, and their associated bond paths, entitled as molecular graphs (MGs). In other words, the topological properties of electron density, ρ
b, along with its derivatives at critical points (CPs), are defined as useful tool to ascertain the concept of chemical bonding and the nature of interatomic interactions in molecular systems. Based on QTAIM formalism, chemical bonding is characterized by duo-gradient paths of charge density function that emanate from a BCP (where the gradient of charge density vanishes mathematically and the Hessian matrix of charge density has two negative and one positive eigenvalues) and end on the corresponding atomic nuclei. To present the more concise interpretation of the nature of bonding and interactions, Laplacian of electron density,
In this context, we investigated the topological analysis of electron density at some key BCPs and RCPs in regioisomeric products and their corresponded transition states via the cycloaddition of cyclooctyne, methyl- and trifluoromethyl-substituted cyclooctynes, and mono and difluorinated cyclooctynes with methyl azide. In particular, we made a topological comparison of electron density and its derivatives at BCPs and RCPs to discern the electronic aspects of bonding and interactions in the enhancement of reactivity.
We have depicted the complete QTAIM molecular graphs of isomeric products and their corresponded transition states in Figures 7–11, for cycloaddition reaction of methyl azide with cyclooctyne, methyl- and trifluoromethyl-substituted cyclooctynes, and mono and difluorinated cyclooctynes, respectively. The QTAIM molecular graphs have been calculated at M08-HX/6-311+G** level of theory, which consist of all CPs of electron density and their associated bond paths.

Complete MGs of product-HH and TS-HH, obtained by QTAIM analysis of M08-HX/6-311+G** electron density. BCPs: blue circles; RCPs: red circles; and Bond Paths: black lines.

Complete MGs of isomer1-HMe, isomer2-HMe, TS1-HMe, and TS2-HMe, obtained by QTAIM analysis of M08-HX/6-311+G** electron density.

Complete MGs of isomer1-HCF3, isomer2-HCF3, TS1-HCF3, and TS2-HCF3, obtained by QTAIM analysis of M08-HX/6-311+G** electron density.

Complete MGs of isomer1-HF, isomer2-HF, TS1-HF, and TS2-HF, obtained by QTAIM analysis of M08-HX/6-311+G** electron density.

Complete MGs of isomer1-FF, isomer2-FF, TS1-FF, and TS2-FF, obtained by QTAIM analysis of M08-HX/6-311+G** electron density.
In Tables 1–5, we have reported the calculated values of electron density, its Laplacian, kinetic and potential energy densities, and also the total electronic energy density at some key BCPs and RCPs of isomer1 and isomer2 products and their corresponded transition states for cycloaddition reaction of methyl azide with cyclooctyne, methyl- and trifluoromethyl-substituted cyclooctynes, and mono and difluorinated cyclooctynes, respectively.
Mathematical properties of some selected BCPs and RCPs of product-HH and TS-HH
|
|
|
|
|
|
---|---|---|---|---|---|
Product-HH | |||||
BCP S | |||||
N(2)–C(5) | 0.280 | −0.578 | 0.231 | −0.607 | −0.376 |
N(3)–C(4) | 0.297 | −0.799 | 0.174 | −0.548 | −0.374 |
C(4)–C(5) | 0.308 | −0.810 | 0.114 | −0.431 | −0.317 |
H(8)–H(21) | 0.017 | 0.057 | 0.012 | −0.009 | 0.002 |
H(17)–H(20) | 0.019 | 0.068 | 0.015 | −0.012 | 0.002 |
RCPs | |||||
RCP1 | 0.053 | 0.492 | 0.097 | −0.087 | 0.010 |
RCP2 | 0.011 | 0.047 | 0.009 | −0.008 | 0.001 |
RCP3 | 0.011 | 0.053 | 0.011 | −0.008 | 0.003 |
RCP4 | 0.010 | 0.049 | 0.009 | −0.008 | 0.001 |
RCP5 | 0.014 | 0.068 | 0.014 | −0.011 | 0.003 |
TS-HH | |||||
BCPs | |||||
N(9)–C(7) | 0.049 | 0.090 | 0.028 | −0.033 | −0.005 |
N(8)–C(1) | 0.060 | 0.088 | 0.031 | −0.04 | −0.009 |
C(1)–C(7) | 0.399 | −1.236 | 0.238 | −0.806 | −0.557 |
RCPs | |||||
RCP1 | 0.023 | 0.149 | 0.031 | −0.025 | 0.006 |
RCP2 | 0.009 | 0.038 | 0.008 | −0.006 | 0.002 |
Note: These properties are obtained through QTAIM analysis at M08-HX/6-311+G** level of theory. The atoms are numbered according to Figure 7.
Mathematical properties of some selected BCPs and RCPs of isomer1-HMe, isomer2-HMe, TS1-HMe, and TS2-HMe
|
|
|
|
|
|
---|---|---|---|---|---|
Isomer1-HMe | |||||
BCP S | |||||
N(24)–C(23) | 0.278 | −0.566 | 0.231 | −0.603 | −0.372 |
N(30)–C(22) | 0.297 | −0.797 | 0.175 | −0.549 | −0.374 |
C(22)–C(23) | 0.306 | −0.801 | 0.114 | −0.428 | −0.314 |
H(9)–H(17) | 0.009 | 0.030 | 0.006 | −0.005 | 0.001 |
RCPs | |||||
RCP1 | 0.053 | 0.428 | 0.097 | −0.087 | 0.010 |
RCP2 | 0.008 | 0.034 | 0.007 | −0.005 | 0.020 |
RCP3 | 0.008 | 0.034 | 0.007 | −0.006 | 0.001 |
Isomer2-HMe | |||||
BCP S | |||||
N(25)–C(2) | 0.277 | −0.567 | 0.227 | −0.597 | −0.370 |
N(26)–C(1) | 0.297 | −0.798 | 0.172 | −0.543 | −0.371 |
N(25)–C(27) | 0.246 | −0.573 | 0.154 | −0.451 | −0.297 |
C(1)–C(2) | 0.307 | −0.805 | 0.115 | −0.431 | −0.316 |
H(9)–H(28) | 0.005 | 0.017 | 0.004 | −0.003 | 0.001 |
H(11)–H(21) | 0.018 | 0.057 | 0.012 | −0.010 | 0.002 |
RCP S | |||||
RCP1 | 0.053 | 0.428 | 0.097 | −0.087 | 0.010 |
RCP2 | 0.008 | 0.031 | 0.006 | −0.005 | 0.001 |
RCP3 | 0.008 | 0.031 | 0.007 | −0.005 | 0.002 |
RCP4 | 0.005 | 0.019 | 0.004 | −0.003 | 0.001 |
TS1-HMe | |||||
BCP S | |||||
N(24)–C(1) | 0.044 | 0.088 | 0.026 | −0.029 | −0.003 |
N(26)–C(2) | 0.058 | 0.089 | 0.031 | −0.039 | −0.008 |
C(1)–C(2) | 0.399 | −1.239 | 0.249 | −0.808 | −0.559 |
RCP S | |||||
RCP1 | 0.022 | 0.141 | 0.029 | −0.023 | 0.006 |
RCP2 | 0.008 | 0.034 | 0.007 | −0.005 | 0.002 |
TS2-HMe | |||||
BCP S | |||||
N(10)–C(1) | 0.065 | 0.089 | 0.033 | −0.044 | −0.011 |
N(9)–C(2) | 0.046 | 0.088 | 0.026 | −0.030 | −0.004 |
C(1)–C(2) | 0.398 | −1.238 | 0.246 | −0.801 | −0.555 |
RCP S | |||||
RCP1 | 0.022 | 0.148 | 0.031 | −0.025 | 0.006 |
RCP2 | 0.009 | 0.038 | 0.008 | −0.006 | 0.002 |
Note: These properties are obtained through QTAIM analysis at M08-HX/6-311+G** level of theory. The atoms are numbered according to Figure 8.
Mathematical properties of some selected BCPs and RCPs of isomer1-HCF3, isomer2-HCF3, TS1-HCF3, and TS2-HCF3
|
|
|
|
|
|
---|---|---|---|---|---|
Isomer1-HCF 3 | |||||
BCP S | |||||
N(14)–C(2) | 0.296 | −0.791 | 0.166 | −0.53 | −0.364 |
N(13)–C(1) | 0.279 | −0.594 | 0.225 | −0.599 | −0.374 |
C(1)–C(2) | 0.305 | −0.797 | 0.114 | −0.427 | −0.313 |
H(26)–F(10) | 0.031 | 0.143 | 0.033 | −0.030 | 0.003 |
H(30)–F(11) | 0.009 | 0.039 | 0.008 | −0.007 | 0.001 |
H(30)–F(10) | 0.009 | 0.039 | 0.009 | −0.008 | 0.001 |
RCPs | |||||
RCP1 | 0.053 | 0.427 | 0.097 | −0.087 | 0.010 |
RCP2 | 0.007 | 0.031 | 0.006 | −0.005 | 0.001 |
RCP3 | 0.007 | 0.028 | 0.006 | −0.005 | 0.001 |
RCP4 | 0.008 | 0.035 | 0.008 | −0.006 | 0.002 |
RCP5 | 0.009 | 0.044 | 0.009 | −0.008 | 0.001 |
RCP6 | 0.009 | 0.042 | 0.009 | −0.007 | 0.002 |
Isomer2-HCF 3 | |||||
BCP S | |||||
N(14)–C(1) | 0.279 | −0.609 | 0.215 | −0.582 | −0.367 |
N(16)–C(2) | 0.299 | −0.815 | 0.174 | −0.552 | −0.378 |
C(1)–C(2) | 0.304 | −0.788 | 0.113 | −0.423 | −0.310 |
H(30)–F(11) | 0.032 | −0.147 | 0.034 | −0.031 | 0.003 |
H(7)–H(22) | 0.020 | 0.068 | 0.015 | −0.013 | 0.002 |
RCP S | |||||
RCP1 | 0.053 | 0.428 | 0.097 | −0.087 | 0.010 |
RCP2 | 0.009 | 0.050 | 0.010 | −0.008 | 0.002 |
RCP3 | 0.011 | 0.054 | 0.011 | −0.009 | 0.002 |
RCP4 | 0.008 | 0.041 | 0.008 | −0.006 | 0.002 |
TS1-HCF 3 | |||||
BCP S | |||||
N(14)–C(1) | 0.066 | 0.092 | 0.036 | −0.049 | −0.013 |
N(13)–C(2) | 0.047 | 0.081 | 0.024 | −0.028 | −0.004 |
C(1)–C(2) | 0.394 | −1.206 | 0.244 | −0.791 | −0.547 |
N(13)–F(12) | 0.010 | 0.045 | 0.009 | −0.008 | 0.001 |
RCP S | |||||
RCP1 | 0.023 | 0.154 | 0.032 | −0.026 | 0.006 |
RCP2 | 0.008 | 0.033 | 0.007 | −0.005 | 0.002 |
RCP3 | 0.009 | 0.039 | 0.008 | −0.007 | 0.001 |
TS2-HCF 3 | |||||
BCP S | |||||
N(10)–C(1) | 0.053 | 0.089 | 0.029 | −0.036 | −0.007 |
N(9)–C(2) | 0.062 | 0.088 | 0.032 | −0.042 | −0.010 |
C(1)–C(2) | 0.399 | −1.239 | 0.244 | −0.799 | −0.555 |
N(10)–F(19) | 0.008 | 0.033 | 0.007 | −0.006 | 0.001 |
F(19)–H(13) | 0.008 | 0.037 | 0.008 | −0.006 | 0.002 |
RCP S | |||||
RCP1 | 0.024 | 0.159 | 0.033 | −0.027 | 0.006 |
RCP2 | 0.009 | 0.039 | 0.008 | −0.006 | 0.002 |
RCP3 | 0.007 | 0.037 | 0.007 | −0.006 | 0.001 |
RCP4 | 0.008 | 0.033 | 0.007 | −0.006 | 0.001 |
Note: The atoms are numbered according to Figure 9.
Mathematical properties of some selected BCPs and RCPs of isomer1-HF, isomer2-HF, TS1-HF, and TS2-HF
|
|
|
|
|
|
---|---|---|---|---|---|
Isomer1-HF | |||||
BCP S | |||||
N(9)–C(2) | 0.299 | −0.805 | 0.170 | −0.542 | −0.372 |
N(10)–C(1) | 0.280 | −0.588 | 0.227 | −0.602 | −0.375 |
C(2)–C(1) | 0.305 | −0.798 | 0.114 | −0.428 | −0.314 |
H(12)–H(21) | 0.007 | 0.025 | 0.005 | −0.004 | 0.0010 |
RCPs | |||||
RCP1 | 0.053 | 0.429 | 0.097 | −0.087 | 0.010 |
RCP2 | 0.007 | 0.030 | 0.006 | −0.005 | 0.001 |
RCP3 | 0.006 | 0.028 | 0.006 | −0.005 | 0.001 |
RCP4 | 0.007 | 0.027 | 0.006 | −0.004 | 0.002 |
Isomer2-HF | |||||
BCP S | |||||
N(22)–C(2) | 0.280 | −0.581 | 0.230 | −0.605 | −0.375 |
N(21)–C(1) | 0.300 | −0.816 | 0.173 | −0.551 | −0.378 |
C(2)–C(1) | 0.306 | −0.801 | 0.114 | −0.428 | −0.314 |
H(12)–H(19) | 0.019 | −0.064 | 0.014 | −0.012 | 0.0020 |
RCP S | |||||
RCP1 | 0.054 | 0.430 | 0.097 | −0.087 | 0.010 |
RCP2 | 0.010 | 0.050 | 0.010 | −0.008 | 0.002 |
RCP3 | 0.011 | 0.053 | 0.011 | −0.009 | 0.002 |
RCP4 | 0.009 | 0.043 | 0.009 | −0.007 | 0.002 |
TS1-HF | |||||
BCP S | |||||
N(9)–C(2) | 0.051 | 0.086 | 0.026 | −0.032 | −0.005 |
N(10)–C(1) | 0.061 | 0.093 | 0.034 | −0.044 | −0.010 |
C(2)–C(1) | 0.397 | −1.218 | 0.247 | −0.799 | −0.552 |
RCP S | |||||
RCP1 | 0.023 | 0.156 | 0.033 | −0.023 | 0.007 |
RCP2 | 0.009 | 0.035 | 0.007 | −0.006 | 0.001 |
TS2-HF | |||||
BCP S | |||||
N(21)–C(2) | 0.060 | 0.090 | 0.032 | −0.042 | −0.010 |
N(22)–C(1) | 0.054 | 0.085 | 0.028 | −0.034 | −0.006 |
C(2)–C(1) | 0.398 | −1.231 | 0.243 | −0.793 | −0.550 |
H(25)–F(20) | 0.010 | 0.042 | 0.009 | −0.008 | 0.001 |
RCP S | |||||
RCP1 | 0.024 | 0.161 | 0.034 | −0.027 | 0.007 |
RCP2 | 0.009 | 0.036 | 0.007 | −0.006 | 0.001 |
RCP3 | 0.008 | 0.039 | 0.008 | −0.007 | 0.001 |
Note: The atoms are numbered according to Figure 10.
Mathematical properties of some selected BCPs and RCPs of isomer1-FF, isomer2-FF, TS1-FF, and TS2-FF
|
|
|
|
|
|
---|---|---|---|---|---|
Isomer1-FF | |||||
BCP S | |||||
N(21)–C(2) | 0.331 | −0.886 | 0.248 | −0.717 | −0.469 |
N(22)–C(1) | 0.309 | −0.452 | 0.338 | −0.789 | −0.451 |
C(2)–C(1) | 0.311 | −0.832 | 0.117 | −0.442 | −0.325 |
RCPs | |||||
RCP1 | 0.059 | 0.474 | 0.109 | −0.10 | 0.009 |
RCP2 | 0.006 | 0.027 | 0.005 | −0.004 | 0.001 |
RCP3 | 0.006 | 0.028 | 0.006 | −0.004 | 0.002 |
RCP4 | 0.007 | 0.033 | 0.007 | −0.006 | 0.001 |
Isomer2-FF | |||||
BCP S | |||||
N(22)–C(2) | 0.281 | −0.593 | 0.227 | −0.602 | −0.375 |
N(21)–C(1) | 0.302 | −0.828 | 0.172 | −0.552 | −0.379 |
C(2)–C(1) | 0.306 | −0.796 | 0.114 | −0.426 | −0.313 |
H(21)–H(19) | 0.019 | −0.064 | 0.014 | −0.012 | 0.002 |
RCP S | |||||
RCP1 | 0.054 | 0.431 | 0.098 | −0.087 | 0.010 |
RCP2 | 0.010 | 0.049 | 0.010 | −0.008 | 0.002 |
RCP3 | 0.011 | 0.053 | 0.011 | −0.009 | 0.003 |
RCP4 | 0.008 | 0.041 | 0.009 | −0.007 | 0.002 |
TS1-FF | |||||
BCP S | |||||
N(22)–C(1) | 0.064 | 0.094 | 0.035 | −0.046 | −0.011 |
N(21)–C(2) | 0.050 | 0.086 | 0.026 | −0.031 | −0.005 |
C(2)–C(1) | 0.398 | −1.222 | 0.251 | −0.808 | −0.557 |
RCP S | |||||
RCP1 | 0.024 | 0.159 | 0.033 | −0.027 | 0.006 |
RCP2 | 0.009 | 0.035 | 0.007 | −0.005 | 0.002 |
TS2-FF | |||||
BCP S | |||||
N(9)–C(1) | 0.049 | 0.083 | 0.025 | −0.030 | −0.005 |
N(8)–C(2) | 0.066 | 0.094 | 0.036 | −0.048 | −0.012 |
C(2)–C(1) | 0.399 | −1.233 | 0.245 | −0.799 | −0.554 |
H(12)–F(16) | 0.007 | 0.030 | 0.006 | −0.005 | 0.001 |
RCP S | |||||
RCP1 | 0.024 | 0.160 | 0.034 | −0.027 | 0.007 |
RCP2 | 0.009 | 0.037 | 0.007 | −0.006 | 0.001 |
RCP3 | 0.007 | 0.032 | 0.007 | −0.005 | 0.002 |
Note: The atoms are numbered according to Figure 11.
The comparative assessment on the calculated QTAIM results of BCPs comprehensibly ascertains the following facts: (1) In all products and their corresponded transition states, the negative values of
On the other hand, the more stringent analysis of RCPs on QTAIM MGs of regioisomeric products and their corresponded transition states (Figures 7–11) demonstrates the increased electronic stability of products in comparison with transition states from electron density topological viewpoints. Clearly speaking, due to the presence of H–H, H–F, and F–C intramolecular semi-covalent semi-electrostatic interactions in regioisomeric products (with the small and positive
4 Conclusion
In this research, the effect of substitution at the propargylic position in cyclooctyne on the enhancement of reactivity in the strain-promoted azide–alkyne cycloadditions was assessed via the quantum chemistry calculations. We demonstrated how the QTAIM approach provides significant insight into the activity enhancement of strain-promoted azide–alkyne cycloadditions by methyl, trifluoromethyl, and fluorine substitution of propargylic position in cyclooctyne. In this respect, we primarily surveyed the energetic aspects in the mechanical pathway of cycloaddition reactions, considering the regioisomeric transition state structures and interpret the regioselectivity of reaction and methyl, trifluoromethyl, and fluorine substituent effects on the reactivity.
On the other hand, topological analysis of electron density was performed on some key BCPs and RCPs in regioisomeric products and their corresponded transition states along the cycloaddition of cyclooctyne, methyl-and trifluoromethyl-substituted cyclooctyne, and mono and difluorinated cyclooctynes with methyl azide. Based on the QTAIM calculated results, we have arrived at this insight that the enhanced cycloaddition reactivity of trifluoromethyl-substituted and fluorinated cycloalkynes can be substantially originated from the stabilizing topological properties of electron density and its derivatives upon trifluoromethyl substitution and fluorination and not just from the energetic aspects along the cycloaddition pathway.
In summary, our quantum chemistry investigation clearly proved that the promoted cycloaddition reaction of methyl azide with trifluoromethyl-substituted and fluorinated cyclooctynes benefited from both reduced activation energies and stabilized topological features of electron density.
Acknowledgements
The authors are grateful to the research council of Alzahra University for all the support received during the research process.
-
Funding information: The authors gratefully acknowledge the partial financial support received from the research council of Alzahra University.
-
Conflict of interest: The authors state no conflict of interest.
-
Data availability statement: The datasets generated during and/or analyzed during the current study are available from the corresponding author on reasonable request.
References
[1] Agard NJ, Prescher JA, Bertozzi CR. A strain-promoted [3+2] azide−alkyne cycloaddition for covalent modification of biomolecules in living system. J Am Chem Soc. 2004;126(46):15046–7.10.1021/ja044996fSearch in Google Scholar PubMed
[2] Lutz JF. Copper‐free azide–alkyne cycloadditions: new insights and perspectives. Angew Chem Int Ed. 2008;47(12):2182–4.10.1002/anie.200705365Search in Google Scholar PubMed
[3] Ess DH, Jones GO, Houk KN. Transition states of strain-promoted metal-free click chemistry: 1,3-dipolar cycloadditions of phenyl azide and cyclooctynes. Org Lett. 2008;10(8):1633–6.10.1021/ol8003657Search in Google Scholar PubMed
[4] Levandowski BJ, Gamache RF, Murphy JM, Houk KN. Readily accessible ambiphilic cyclopentadienes for bioorthogonal labeling. J Am Chem Soc. 2018;140(20):6426–31.10.1021/jacs.8b02978Search in Google Scholar PubMed PubMed Central
[5] King M, Wagner A. Developments in the field of bioorthogonal bond forming reactions-past and present trends. Bioconjugate Chem. 2014;25(5):825–39.10.1021/bc500028dSearch in Google Scholar PubMed
[6] Debets MF, van der Doelen CWJ, Rutjes FPJT, van Delft FL. Azide: a unique dipole for metal-free bioorthogonal ligations. Chem Bio Chem. 2010;11(9):1168–84.10.1002/cbic.201000064Search in Google Scholar PubMed
[7] Saxon E. Cell surface engineering by a modified staudinger reaction. Science. 2000;287(5460):2007–10.10.1126/science.287.5460.2007Search in Google Scholar PubMed
[8] Lang K, Chin JW. Cellular incorporation of unnatural amino acids and bioorthogonal labeling of proteins. Chem Rev. 2014;114(9):4764–806.10.1021/cr400355wSearch in Google Scholar PubMed
[9] Shieh P, Bertozzi CR. Design strategies for bioorthogonal smart probes. Org Biomol Chem. 2014;12(46):9307–20.10.1039/C4OB01632GSearch in Google Scholar
[10] Lang K, Chin JW. Bioorthogonal reactions for labeling proteins. ACS Chem Biol. 2014;9(1):16–20.10.1021/cb4009292Search in Google Scholar PubMed
[11] Liu F, Liang Y, Houk KN. Bioorthogonal cycloadditions: computational analysis with the distortion/interaction model and predictions of reactivities. Acc Chem Res. 2017;50(9):2297–308.10.1021/acs.accounts.7b00265Search in Google Scholar
[12] Liang Y, Mackey JL, Lopez SA, Liu F, Houk KN. Control and design of mutual orthogonality in bioorthogonal cycloadditions. J Am Chem Soc. 2012;134(43):17904–7.10.1021/ja309241eSearch in Google Scholar
[13] Agard NJ, Baskin JM, Prescher JA, Lo A, Bertozzi CR. A comparative study of bioorthogonal reactions with azides. ACS Chem Biol. 2006;1(10):644–8.10.1021/cb6003228Search in Google Scholar
[14] Nguyen LT, De Proft F, Dao VL, Ngyuen MT, Geerlings PJ. A theoretical approach to the regioselectivity in 1,3‐dipolar cycloadditions of diazoalkanes, hydrazoic acid and nitrous oxide to acetylenes, phosphaalkynes and cyanides. Phys Org Chem. 2003;16(9):615–25.10.1002/poc.653Search in Google Scholar
[15] Gordon CG, Mackey JL, Jewett JC, Sletten EM, Houk KN, Bertozzi CR. Reactivity of biarylazacyclooctynones in copper-free click chemistry. J Am Chem Soc. 2012;134(22):9199–208.10.1021/ja3000936Search in Google Scholar
[16] Ni R, Mitsuda N, Kashiwagi T, Igawa K, Tomooka K. Heteroatom‐embedded medium‐sized cycloalkynes: concise synthesis, structural analysis, and reactions. Angew Chem Int Ed. 2015;54(4):1190–4.10.1002/anie.201409910Search in Google Scholar
[17] Rostovtsev VV, Green LG, Fokin VV, Sharpless KB. A stepwise Huisgen cycloaddition process: copper(i)‐catalyzed regioselective “ligation” of azides and terminal alkynes. Chem Int Ed. 2002;41(14):2596–9.10.1002/1521-3773(20020715)41:14<2596::AID-ANIE2596>3.0.CO;2-4Search in Google Scholar
[18] Tornoe CW, Christensen C, Meldal M. Peptidotriazoles on solid phase: [1,2,3]-triazoles by regiospecific copper(i)-catalyzed 1,3-dipolar cycloadditions of terminal alkynes to azides. J Org Chem. 2002;67(9):3057–64.10.1021/jo011148jSearch in Google Scholar
[19] Link AJ, Tirrell DA. Cell surface labeling of escherichia coli via copper(i)-catalyzed [3+2] cycloaddition. J Am Chem Soc. 2003;125(37):11164–5.10.1021/ja036765zSearch in Google Scholar
[20] Lallana E, Fernandez-Megia E, Riguera R. Surpassing the use of copper in the click functionalization of polymeric nanostructures: a strain-promoted approach. J Am Chem Soc. 2009;131(16):5748–50.10.1021/ja8100243Search in Google Scholar
[21] Laughlin ST, Baskin JM, Amacher SL, Bertozzi CR. In vivo imaging of membrane-associated glycans in developing zebrafish. Science. 2008;320(5876):664–7.10.1126/science.1155106Search in Google Scholar PubMed PubMed Central
[22] Shih HW, Kamber DN, Prescher JA. Building better bioorthogonal reactions. Curr Opin Chem Biol. 2014;21:103–11.10.1016/j.cbpa.2014.07.002Search in Google Scholar PubMed
[23] Vocadlo DJ, Hang HC, Kim EJ, Hanover JA, Bertozzi CR. A metabolic labeling approach toward proteomic analysis of mucin-type O-linked glycosylation. Proc Natl Acad Sci. 2003;100(25):14846–51.10.1073/pnas.2335201100Search in Google Scholar PubMed PubMed Central
[24] Selvaraj R, Fox JM. Trans-cyclooctene: a stable, voracious dienophile for bioorthogonal labeling. Curr Opin Chem Biol. 2013;17(5):753–60.10.1016/j.cbpa.2013.07.031Search in Google Scholar PubMed PubMed Central
[25] Dommerholt J, Schmidt S, Temming R, Hendriks LJA, Rutjes FPJT, van Hest JCM, et al. Readily accessible bicyclononynes for bioorthogonal labeling and three-dimensional imaging of living cells. Angew Chem Int Ed. 2010;49(49):9422–5.10.1002/anie.201003761Search in Google Scholar PubMed PubMed Central
[26] Debets MF, van Berkel SS, Schoffelen S, Rutjes FPJT, van Hest JCM, van Delft FL. Aza-dibenzocyclooctynes for fast and efficient enzyme PEGylation via copper-free (3+2) cycloaddition. Chem Commun. 2010;46(1):97–9.10.1039/B917797CSearch in Google Scholar PubMed
[27] Sletten EM, Nakamura H, Jewett JC, Bertozzi CR. Difluorobenzocyclooctyne: synthesis, reactivity, and stabilization by β-cyclodextrin. J Am Chem Soc. 2010;132(33):11799–805.10.1021/ja105005tSearch in Google Scholar PubMed PubMed Central
[28] Ning X, Guo J, Wolfert MA, Boons GJ. Visualizing metabolically labeled glycoconjugates of living cells by copper-free and fast Huisgen cycloadditions. Angew Chem Int Ed. 2008;47(12):2253–5.10.1002/anie.200705456Search in Google Scholar PubMed PubMed Central
[29] Ess DH, Houk KN. Distortion/Interaction energy control of 1,3-dipolar cycloaddition reactivity. J Am Chem Soc. 2007;129(35):10646–7.10.1021/ja0734086Search in Google Scholar PubMed
[30] Ess DH, Houk KN. Theory of 1,3-dipolar cycloadditions: distortion/interaction and frontier molecular orbital models. J Am Chem Soc. 2008;130(31):10187–98.10.1021/ja800009zSearch in Google Scholar PubMed
[31] Hamlin TA, Levandowski BJ, Narsaria AK, Houk KN, Matthias, Bickelhaupt F. Structural distortion of cycloalkynes influences cycloaddition rates both by strain and interaction energies. Chem Eur J. 2019;25(25):6342–8.10.1002/chem.201900295Search in Google Scholar PubMed PubMed Central
[32] Becke AD. A new mixing of Hartree–Fock and local density‐functional theories. J Chem Phys. 1993;98(2):1372–7.10.1063/1.464304Search in Google Scholar
[33] Hosseinnejad T, Heravi MM, Firouzi R. Regioselectivity in the sonogashira synthesis of 6-(4-nitrobenzyl)-2-phenylthiazolo[3,2-b]1,2,4 triazole: a quantum chemistry study. J Mol Modeling. 2013;19(2):951–61.10.1007/s00894-012-1639-1Search in Google Scholar PubMed
[34] Hosseinnejad T, Mahdavian S. Quantum chemistry study on regioselectivity in ruthenium catalyzed synthesis of 1,5-disubstituted 1,2,3-triazoles. Comp Theo Chem. 2018;1143:29–35.10.1016/j.comptc.2018.09.002Search in Google Scholar
[35] Bader RFW. Atoms in molecules: a quantum theory. Oxford, UK: Clarendon Press; 1990.10.1093/oso/9780198551683.001.0001Search in Google Scholar
[36] Bader RFW, Essén H. The characterization of atomic interactions. J Chem Phys. 1984;80(5):1943–60.10.1063/1.446956Search in Google Scholar
[37] Hosseinnejad T, Dinyari M. Computational study on stereoselective synthesis of substituted 1H-tetrazolesvia a click reaction: DFT and QTAIM approaches. Comput Theor Chem. 2015;1071:53–60.10.1016/j.comptc.2015.08.014Search in Google Scholar
[38] Huisgen R. Kinetics and reaction mechanisms. Pure Appl Chem. 1989;61(4):613–28.10.1351/pac198961040613Search in Google Scholar
[39] Heravi MM, Tamimi M, Yahyavi H, Hosseinnejad T. Huisgen’s cycloaddition reactions: a full perspective. Curr Org Chem. 2016;20(15):1591–647.10.2174/1385272820666151217183010Search in Google Scholar
[40] Zhao Y, Truhlar DG. Exploring the limit of accuracy of the global hybrid meta density functional for main-group thermochemistry, kinetics, and noncovalent interactions. J Chem Theor Comp. 2008;4(11):1849–68.10.1021/ct800246vSearch in Google Scholar PubMed
[41] Schmidt MW, Baldridge KK, Boatz JA, Elbert ST, Gordon MS, Jensen JH, et al. General atomic and molecular electronic structure system. J Comput Chem. 1993;14(11):1347–63.10.1002/jcc.540141112Search in Google Scholar
[42] Bader RFW. AIM2000 program ver 2.0. Hamilton: McMaster University; 2000.Search in Google Scholar
[43] Popelier PLA. The chemical bond: fundamental aspects of chemical bonding. In: Frenking G, Shaik S, editors. The chemical bond: fundamental aspects of chemical bonding. Weinheim. Germany: Wiley‐VCH Verlag GmbH & Co; 2014.10.1002/9783527664696.ch8Search in Google Scholar
© 2021 Tayebeh Hosseinnejad and Marzieh Omrani-Pachin, published by De Gruyter
This work is licensed under the Creative Commons Attribution 4.0 International License.
Articles in the same Issue
- Research Articles
- Molecular, Electronic, Nonlinear Optical and Spectroscopic Analysis of Heterocyclic 3-Substituted-4-(3-methyl-2-thienylmethyleneamino)-4,5-dihydro-1H-1,2,4-triazol-5-ones: Experiment and DFT Calculations
- Lewis acid / Base-free Strategy for the Synthesis of 2-Arylthio and Selenyl Benzothiazole / Thiazole and Imidazole
- Synergistic promoting effect of ball milling and Fe(ii) catalysis for cross-dehydrogenative-coupling of 1,4-benzoxazinones with indoles
- Magnetic nanoparticle-supported sulfonic acid as a green catalyst for the one-pot synthesis of 2,4,5-trisubstituted imidazoles and 1,2,4,5-tetrasubstituted imidazoles under solvent-free conditions
- Synthesis, characteristic fragmentation patterns, and antibacterial activity of new azo compounds from the coupling reaction of diazobenzothiazole ions and acetaminophen
- Para toluenesulfonic acid-catalyzed one-pot, three-component synthesis of benzo[5,6]chromeno[3,2-c]quinoline compounds in aqueous medium
- Design and microwave-assisted synthesis of a novel Mannich base and conazole derivatives and their biological assessment
- Quantum chemical studies on structural, spectroscopic, nonlinear optical, and thermodynamic properties of the 1,2,4-triazole compound
- Design, synthesis, and biological evaluation of phenyl-isoxazole-carboxamide derivatives as anticancer agents
- Quantum chemistry study on the promoted reactivity of substituted cyclooctynes in bioorthogonal cycloaddition reactions
- Quantum chemical calculations of phenazine-based organic dyes in dye-sensitized solar cells
- Rapid Communications
- One-pot synthesis of benzofurans via heteroannulation of benzoquinones
- Design and development of novel thiazole-sulfonamide derivatives as a protective agent against diabetic cataract in Wistar rats via inhibition of aldose reductase
- Review Articles
- Organic electrochemistry: Synthesis and functionalization of β-lactams in the twenty-first century
- Some applications of deep eutectic solvents in alkylation of heterocyclic compounds: A review of the past 10 years
Articles in the same Issue
- Research Articles
- Molecular, Electronic, Nonlinear Optical and Spectroscopic Analysis of Heterocyclic 3-Substituted-4-(3-methyl-2-thienylmethyleneamino)-4,5-dihydro-1H-1,2,4-triazol-5-ones: Experiment and DFT Calculations
- Lewis acid / Base-free Strategy for the Synthesis of 2-Arylthio and Selenyl Benzothiazole / Thiazole and Imidazole
- Synergistic promoting effect of ball milling and Fe(ii) catalysis for cross-dehydrogenative-coupling of 1,4-benzoxazinones with indoles
- Magnetic nanoparticle-supported sulfonic acid as a green catalyst for the one-pot synthesis of 2,4,5-trisubstituted imidazoles and 1,2,4,5-tetrasubstituted imidazoles under solvent-free conditions
- Synthesis, characteristic fragmentation patterns, and antibacterial activity of new azo compounds from the coupling reaction of diazobenzothiazole ions and acetaminophen
- Para toluenesulfonic acid-catalyzed one-pot, three-component synthesis of benzo[5,6]chromeno[3,2-c]quinoline compounds in aqueous medium
- Design and microwave-assisted synthesis of a novel Mannich base and conazole derivatives and their biological assessment
- Quantum chemical studies on structural, spectroscopic, nonlinear optical, and thermodynamic properties of the 1,2,4-triazole compound
- Design, synthesis, and biological evaluation of phenyl-isoxazole-carboxamide derivatives as anticancer agents
- Quantum chemistry study on the promoted reactivity of substituted cyclooctynes in bioorthogonal cycloaddition reactions
- Quantum chemical calculations of phenazine-based organic dyes in dye-sensitized solar cells
- Rapid Communications
- One-pot synthesis of benzofurans via heteroannulation of benzoquinones
- Design and development of novel thiazole-sulfonamide derivatives as a protective agent against diabetic cataract in Wistar rats via inhibition of aldose reductase
- Review Articles
- Organic electrochemistry: Synthesis and functionalization of β-lactams in the twenty-first century
- Some applications of deep eutectic solvents in alkylation of heterocyclic compounds: A review of the past 10 years