Introducing new reactivity descriptors: “Bond reactivity indices.” Comparison of the new definitions and atomic reactivity indices Jesús Sánchez-Márquez Citation: The Journal of Chemical Physics 145, 194105 (2016); doi: 10.1063/1.4967293 View online: http://dx.doi.org/10.1063/1.4967293 View Table of Contents: http://scitation.aip.org/content/aip/journal/jcp/145/19?ver=pdfcov Published by the AIP Publishing Articles you may be interested in Fundamental aspects of recoupled pair bonds. I. Recoupled pair bonds in carbon and sulfur monofluoride J. Chem. Phys. 142, 034113 (2015); 10.1063/1.4905271 A comprehensive analysis of molecule-intrinsic quasi-atomic, bonding, and correlating orbitals. I. HartreeFock wave functions J. Chem. Phys. 139, 234107 (2013); 10.1063/1.4840776 A comparative study of Hamilton and overlap population methods for the analysis of chemical bonding J. Chem. Phys. 113, 1698 (2000); 10.1063/1.481971 A new interpretation of the bonding and spectroscopy of the tetraoxoferrate(VI) FeO 4 2− ion J. Chem. Phys. 109, 6396 (1998); 10.1063/1.477283 Measuring orbitals and bonding in atoms, molecules, and solids Am. J. Phys. 65, 544 (1997); 10.1119/1.18586 Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 150.214.4.235 On: Mon, 05 Dec 2016 07:54:38 THE JOURNAL OF CHEMICAL PHYSICS 145, 194105 (2016) Introducing new reactivity descriptors: “Bond reactivity indices.” Comparison of the new definitions and atomic reactivity indices Jesús Sánchez-Márqueza) Departamento de Quı́mica-Fı́sica, Facultad de Ciencias, Campus Universitario Rı́o San Pedro, Universidad de Cádiz, Puerto Real, Cádiz 11510, Spain (Received 8 June 2016; accepted 26 October 2016; published online 17 November 2016) A new methodology to obtain reactivity indices has been defined. This is based on reactivity functions such as the Fukui function or the dual descriptor and makes it possible to project the information of reactivity functions over molecular orbitals instead of the atoms of the molecule (atomic reactivity indices). The methodology focuses on the molecule’s natural bond orbitals (bond reactivity indices) because these orbitals (with physical meaning) have the advantage of being very localized, allowing the reaction site of an electrophile or nucleophile to be determined within a very precise molecular region. This methodology gives a reactivity index for every Natural Bond Orbital (NBO), and we have verified that they have equivalent information to the reactivity functions. A representative set of molecules has been used to test the new definitions. Also, the bond reactivity index has been related with the atomic reactivity one, and complementary information has been obtained from the comparison. Finally, a new atomic reactivity index has been defined and compared with previous definitions. Published by AIP Publishing. [http://dx.doi.org/10.1063/1.4967293] I. INTRODUCTION Many publications justify the reactivity of a system using canonical orbitals, which sometimes differ from frontier molecular orbitals (HOMO-1, HOMO-2, etc.).1–3 Also, the definition of reactivity index may be ambiguous, as we can see in Ref. 4. This led to a previous study5 to test new definitions for reactivity indices in which we defined Natural Bond Orbital (NBO) reactivity indices and studied their advantages and disadvantages. NBOs6,7 were chosen for two reasons: first, they are very localized and allow the point of attack of an electrophile or nucleophile inside a concrete region of the atom to be delimited and second, because of the ease of use of Lewis structure diagrams. For example, consider the reaction, Calculating the bond reactivity indices in such a case (NBO1 and NBO2 ) would be of great interest. In the current study, a new methodology has been defined for obtaining reactivity indices for the NBOs (bond reactivity indices) that is based on an adequate least square fitting (Lagrange’s undetermined multipliers8 ), which allows the projection of the information of reactivity functions (Fukui function and the dual descriptor) over molecular orbitals instead of the atoms of the molecule (notice that the bond index concept is not a new idea, as can be seen in the study by Bultinck et al.9 ). These indices have the same information as reactivity functions but are easier to use. II. THEORETICAL BACKGROUND where “Nu:” is a nucleophile with one or more occupied NBOs (lone pairs, double-triple bonds, etc.) and “E” is an electrophile with one or more unoccupied NBOs. Imagine that the reaction can follow two paths: a) Author to whom correspondence should be addressed. Electronic mail: [email protected] 0021-9606/2016/145(19)/194105/12/$30.00 To understand detailed reaction mechanisms such as regio-selectivity requires, in addition to global descriptors,10–18 local reactivity parameters to differentiate the reactive behavior of the atoms forming a molecule. The Fukui function19 (f (r)) and the softness20 (s(r)) are two of the most commonly used reactivity descriptors, ! ∂ ρ (r) f (r) = , ∂N ν ! ! ! (1) ∂ ρ (r) ∂ ρ (r) ∂N s (r) = = · = S · f (r) . ∂µ ν ∂N ν ∂ µ ν The Fukui function is associated primarily with the response of the density function of a system to a change in the number of electrons (N) under the constraint of a constant external potential [v(r)]. The Fukui function also represents the response of the chemical potential of a system to a change in external potential. As the chemical potential is a measure of the intrinsic acidic or base strength, and softness incorporates global reactivity, both parameters provide a pair of indices 145, 194105-1 Published by AIP Publishing. Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 150.214.4.235 On: Mon, 05 Dec 2016 07:54:38 194105-2 Jesús Sánchez-Márquez J. Chem. Phys. 145, 194105 (2016) with which to demonstrate, for example, the specific sites of interaction between two reagents. Due to the discontinuity of the electron density with respect to N, finite difference (FD) approximation leads to three types of Fukui function for a system, namely, f + (r) Eq. (2), f – (r) in Eq. (3) and f 0 (r) in Eq. (4) for nucleophilic, electrophilic, and radical attack, respectively. f + (r) is measured by the electron density change following the addition of an electron, and f – (r) by the electron density change upon removal of an electron. f 0 (r) is approximated as the average of both previous terms, defined as follows: f + (r) = ρN0 +1 (r) − ρN0 (r) , for nucleophilic attack, (2) f − (r) = ρN0 (r) − ρN0 −1 (r) , for electrophilic attack, (3) f 0 (r) = 1 ρN0 +1 (r) − ρN0 −1 (r) , 2 for neutral (or radical) attack. (4) where Cν α are the molecular frontier orbital coefficients and S χν are the atomic orbital overlap matrix elements. The subscripts “H” and “L” are referenced to the HOMO and LUMO orbitals. This definition of the condensed Fukui function Eqs. (9)–(12) has been pioneered by Pérez and Chamorro23,24 and has been used in a variety of studies yielding reliable results.25–27 B. Reactivity indices of natural bond orbitals In a previous study,5 we proposed the reactivity index in the following equation for NBOi : X (13) FFiNBO = |Ci α | 2 fk(NBO) i that is based on the approximation, X fkα ≈ , |Ciα | 2 fk(NBO) i (5) where φα (r) is a particular frontier molecular orbital (FMO) chosen depending upon the value of α = + or α = −. Eq. (5) can be used to develop an approximated definition of the Fukui function, f − (r) = ρHOMO (r) , for electrophilic attack, (6) f + (r) = ρLUMO (r) , for nucleophilic attack, (7) where X X (NBO) 2 ( NBO)∗ (NBO) Cν i + = Cχ i Cν i S χν . χ<ν ν ∈ k (15) The fk(NBO) parameters were calculated with a modified i version of the UCA-FUKUI software28 and the C i coefficients of Eq. (16) were obtained by the least square method by applying Eq. (17), which leads to the linear system of Eq. (18), X φ(HOMO) ≈ Ci φi(NBO) , (16) i φ(HOMO) φ(NBO) dτ j XX (8) Expanding the FMO in terms of the atomic basis functions, the condensed Fukui function at the atom k is X X ∗ α 2 |Cν α | + C χ α Cν α S χν , (9) fk = χ<ν ν ∈k X X 2 ∗ |Cν H | + = C χH Cν H S χν χ<ν ν ∈ k × (electrophilic attack), (10) X X 2 ∗ + C C S |C | χL νL χν ν L χ<ν ν ∈k × (nucleophilic attack), (11) fk+ = ≈ X Ck φ(NBO) φj(NBO) dτ, k (17) Ck Bik Bl j Si l . (18) k i fk− (14) i fk(NBO) i Under frozen orbital approximation (FOA) of Fukui, and neglecting the orbital relaxation effects, the Fukui function can be approximated as 1 ρLUMO (r) − ρHOMO (r) , 2 for neutral (or radical) attack. (12) k A. Frontier molecular orbital approximation f 0 (r) = 1 + fk + fk− (radical attack), 2 FFiNBO The condensation of reactivity descriptors to atoms can be carried out in several ways, such as the famous Yang-Mortier scheme21 or other “Population strategies” as can be seen in Refs. 4 and 22. f α (r) ≈ φα (r)2 , fk0 = l Ai Bl j Si l ≈ XXX k i l The current study presents an improved methodology which can use FMO and FD approximations (the definition of Eq. (13) can only use the FMO approximation). Also, the new methodology, unlike the old one, can be based on other functions such as the dual descriptor (the old methodology was based on the f α (r) function). III. INTRODUCTION TO THE METHODOLOGY Below is a brief summary focused on the basis of the new descriptors fi+ (NBO) and fi− (NBO) . The detailed discussion is presented in section S1 in the supplementary material (ST1 in the supplementary material shows a summary of the nomenclature and definitions used in this study). Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 150.214.4.235 On: Mon, 05 Dec 2016 07:54:38 194105-3 Jesús Sánchez-Márquez J. Chem. Phys. 145, 194105 (2016) If we accept the approximation that NBOs do not change when the molecule loses an electron and only their occupancies change, we get all X orbitals 2 * f − (~r ) ≈ fi− (NBO) φ(NBO) r . (19) i i=1 The ( fi− (NBO) ) set can be estimated by the least square fitting of Eq. (20), *f − (~r ) − , all X orbitals 2 fi− (NBO) i=1 φ(NBO) (~r )2 + d~r = MINIMUM, i - and the linear system29 in Eq. (21) can be derived from Eq. (20), all orbitals X 2 φ(NBO) (~r )2 · φ(NBO) (~r )2 d~r ; j = 1, 2, 3, ..., all orbitals. (NBO) − (NBO) − f (~r ) · φj (~r ) d~r = fi i j (20) (21) i=1 Finally, for the f + (~r ) function, we obtain the system of Eq. (22), all orbitals X 2 φ(NBO) (~r )2 · φ(NBO) (~r )2 d~r ; j = 1, 2, 3, ..., all. (NBO) + (NBO) + f (~r ) · φj (~r ) d~r = fi i j (22) i=1 A. A new descriptor based on a dual descriptor In previous studies,22,30,31 the dual descriptor function (∆f (r)) has been defined as Eq. (23), where ρN (r), ρN+1 (r), and are the densities of the neutral molecule, the anion, and the cation. This function was used in a previous study32 to partition the real space into non-overlapping reactive domains that feature a constant ∆f (r) sign that makes it possible to identify the nucleophilic and electrophilic regions inside a molecule, 2 2 * * (23) ∆f (r) = f + (r) − f − (r) = ρN+1 (r) − 2ρN (r) + ρN−1 (r) ≈ φ(LUMO) r − φ(HOMO) r . By using the approximation Eqs. (S7) and (S13) (see supplementary material), we can write the (f + (r) − f – (r)) function as all X orbitals 2 . ∆f (r) = f + (r) − f − (r) = 2αi − αi− − αi+ · φ(NBO) (r) (24) i ρN−1 (r) i=1 That can be rewritten as ∆f (r) = all X orbitals i=1 2 , ∆fi(NBO) · φ(NBO) (r) i (25) where the parameters ∆fi(NBO) have been obtained by means of the condition, all X orbitals 2 2 (NBO) (NBO) *∆f (r) − ∆fi · φi (~r ) + d~r = MINIMUM. , i=1 This is very similar to Eq. (20). Finally, Eq. (27) can be derived from Eq. (26), ! all orbitals X 2 2 (NBO) 2 (NBO) (NBO) (NBO) φi ∆f (r) · φj (~r ) d~r = ∆fi · (~r ) · φj (~r ) d~r , j = 1, 2, 3, ..., all. (26) (27) i=1 B. Calculation details Previous indices were dimensionless (and this is important) but it is also important that they were normalized: all orbitals all orbitals P P fi− (NBO) = 1, fi+ (NBO) = 1, and ∆fi(NBO) = 0. The method of Lagrange multipliers8 was used to normalize all orbitals P i=1 i=1 i=1 the reactivity indices of Eqs. (20, 22 and 26). Finally, we obtained − *f (~r ) − , all X orbitals i=1 fi− (NBO) 2 all X orbitals φ(NBO) (~r )2 + d~r + λ − (NBO) = MINIMUM, f − 1 i i i=1 (28) Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 150.214.4.235 On: Mon, 05 Dec 2016 07:54:38 194105-4 Jesús Sánchez-Márquez *f + (~r ) − , J. Chem. Phys. 145, 194105 (2016) all X orbitals fi+ (NBO) i=1 2 all X orbitals φ(NBO) (~r )2 + d~r + λ + (NBO) = MINIMUM, f − 1 i i i=1 (29) 2 all X all X orbitals orbitals 2 + d~r + λ (NBO) *∆f (~r ) − ∆fi (NBO) φ(NBO) (~ r ) ∆f (30) = MINIMUM. i i , i=1 i=1 Also, instead of calculating the integrals of Eqs. (28)–(30), we have used a grid of equally spaced points to make the fitting. In all cases, we used 182 points/Å3 and verified that by increasing the amount of points, the results do not change significantly (for example, for the A1 reagent we have used 180 000 points). This approximation converts Eq. (21) into Eq. (31), Eq. (22) into Eq. (32), and Eq. (27) into Eq. (33), X X X 2 φ(NBO) (r )2 · φ(NBO) (r )2 , j = 1, 2, 3, ..., (rm ) = f − (rm ) · φ(NBO) fi− (NBO) (31) m m j i j m X m 2 (rm ) = f + (rm ) · φ(NBO) j X 2 (rk ) = ∆f (rk ) · φ(NBO) j X X k m i fi+ (NBO) X m i ∆fi(NBO) X i k IV. COMPUTATIONAL METHODS All the structures included in this study were optimized at the B3LYP/6-31G(d)33,34 level of theory using the Gaussian09 package.35 For the FMO approximation, the electrophilic Fukui function was evaluated from a single point calculation TABLE I. Sample 1: dienes bearing an electron withdrawing group (EWG) at C1 or/and C2 . Reagent R (C1 ) R’ (C2 ) A1 B1 C1 D1 E1 F1 G1 Cl H Cl H NO2 CN H H Cl Cl CHO H H CN TABLE II. Sample 2: dienes bearing an electron donating group (EDG) at C1 or/and C2 . Reagent R (C1 ) R’ (C2 ) H1 I1 J1 K1 L1 CH3 H CH3 −OCH3 H H CH3 CH3 H −OCH3 TABLE III. Sample 3: dienophiles bearing an EWG. Reagent R A2 B2 C2 D2 Cl CHO NO2 CN φ(NBO) (r )2 · φ(NBO) (r )2 , j = 1, 2, 3, ..., m m i j (32) φ(NBO) (r )2 · φ(NBO) (r )2 , j = 1, 2, 3, .... k k i j (33) in terms of molecular orbital coefficients and overlap integrals. For the FD approximation, the densities were calculated for the neutral molecule, cation and anion from single point calculations (the wave functions were calculated with Gaussian09). The condensed Fukui functions, fk(NBO) Eq. (15), fk− Eq. (9) and fk− , calculated under the FD approximation and atomic populations,36 were obtained using UCA-FUKUI software.28 The fi− (NBO) in Eq. (21), fi+ (NBO) in Eq. (22) and ∆fi (NBO) in Eq. (27) parameters were obtained with a modified version of UCA-FUKUI. V. RESULTS AND DISCUSSION A. Testing the new descriptors fi −(NBO) and f i +(NBO) in a sample of representative molecules: Rationalization of Diels-Alder reactions The interaction between unsymmetrical dienes and dienophiles can give two isomeric adducts, depending upon the relative position of the substituent. The selectivity for the formation of one adduct over the other is called regioselectivity, and this kind of isomer is called a regioisomer. In this class of cycloaddition, the degree of regioselectivity is often high;37–42 yet, it is well established that the more powerful the electron-donating and electron-withdrawing substituents on the diene/dienophile pair, the more regioselective the reaction.3 Regioselectivity has been described in terms of a local hard and soft acid and base (HSAB) principle, and some empirical rules have been proposed to rationalize the TABLE IV. Sample 4: dienophiles bearing an EDG. Reagent R E2 F2 CH3 −OCH3 Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 150.214.4.235 On: Mon, 05 Dec 2016 07:54:38 194105-5 Jesús Sánchez-Márquez J. Chem. Phys. 145, 194105 (2016) TABLE V. fi− (NBO) and fi+ (NBO) parameters in Eqs. (21) and (22) calculated under finite difference (FD) and frontier molecular orbital (FMO) approximations (HOMO and LUMO) for some important NBOs of Fig. 1 (sorted by energy, the highest values are printed in boldface). Partial orbital occupancies have been added in the last column. Schematic representations of NBOs Level NBO Type fi+ (NBO) (FD) fi+ (NBO) (FMO: LUMO) Occupancies 25 24 82 87 BD C1−C4 BD C6−C8 0.116 0.156 0.274 0.315 0.0586 0.1258 fi− (NBO) (FMO: HOMO) Occupancies 0.351 0.442 0.175 -0.003 0.023 1.9424 1.9433 1.9204 1.9755 1.9946 Level NBO Type fi− (NBO) (FD) 23 22 21 20 10 4 9 23 22 21 BD C1−C4 BD C6−C8 LP Cl10 LP Cl10 LP Cl10 0.350 0.399 0.153 0.035 0.074 TABLE VI. fi− (NBO) and fi+ (NBO) in Eqs. (21) and (22) under FD approximation and FMO approximation for some important NBOs of Fig. 2 (sorted by energy). Partial orbital occupancies have been added in the last column. Schematic representations of NBOs Level NBO Type 17 51 BD C1−C2 fi+ (NBO) (FD) fi+ (NBO) (FMO: LUMO) Occupancies 0.0736 0.406 0.730 fi− (NBO) fi− (NBO) Level NBO Type (FD) (FMO: HOMO) Occupancies 16 15 14 8 2 16 15 14 BD C1−C2 LP Cl6 LP Cl6 LP Cl6 0.407 0.212 -0.040 0.031 0.664 0.365 0.003 0.012 1.9970 1.9246 1.9754 1.9946 experimental regioselectivity pattern observed in some DielsAlder (D-A) reactions.43,44 There is not a unique criterion; however, it explains most of the experimental evidence that has accumulated in cycloaddition processes involving fourcenter interactions. An excellent discussion of regioselectivity in concerted pericyclic reactions can be found in Ref. 44. Eqs. (21) and (22) were applied to a sample of molecules that are related to the D-A reaction type (to obtain the fi− (NBO) and fi+ (NBO) parameters). This set of molecules was divided into four samples (Tables I–IV) depending on the substituents (electron withdrawing group (EWG) or electron donating group (EDG)).45 As an example, SF. 1 (supplementary material) shows some NBOs of the reagents A1 and A2 and the schematic representations of these orbitals (this kind of representations will be used in Tables V and VI and Figs. 5–10). FIG. 1. (a) fi− (NBO) and (b) fi+ (NBO) parameters in Eqs. (21) and (22) calculated under frontier molecular orbital approximation for reagent A1 . The NBOs are represented on the “x” axis (sorted by energy). Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 150.214.4.235 On: Mon, 05 Dec 2016 07:54:38 194105-6 Jesús Sánchez-Márquez J. Chem. Phys. 145, 194105 (2016) FIG. 2. (a) fi− (NBO) and (b) fi+ (NBO) parameters in Eqs. (21) and (22) calculated under frontier molecular orbital approximation for reagent A2 . The NBOs are represented on the “x” axis (sorted by energy). FIG. 3. ∆fi (NBO) indices (based on the dual descriptor function) for the reagent A1 . Calculations were made with Eq. (33) and (a) frontier molecular orbital approximation or (b) finite difference approximation. The NBOs are represented on the “x” axis (sorted by energy). Figs. 1 and 2 show parameters fi− (NBO) and fi+ (NBO) of Eqs. (21) and (22) calculated for the reagents A1 and A2 (samples 1 and 3). Tables V and VI show these parameters for some important NBOs of Figs. 1 and 2 (higher values) under finite difference (FD) and frontier molecular orbital (FMO: HOMO and LUMO) approximations. In the supplementary material, FIG. 4. Bond reactivity indices ∆fi (NBO) based on the dual descriptor function for reagent A2 . (a) Frontier molecular orbital approximation, (b) finite difference approximation. The NBOs are represented on the “x” axis (sorted by energy). Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 150.214.4.235 On: Mon, 05 Dec 2016 07:54:38 194105-7 Jesús Sánchez-Márquez J. Chem. Phys. 145, 194105 (2016) TABLE VII. ∆fi (NBO) indices (based on the dual descriptor function) for some important NBOs of Fig. 3 (reagent A1 ). Calculations were made with Eq. (33) and frontier molecular orbital approximation. NBO 4 9 23 22 21 Type ∆fi (NBO) (FMO) ∆fi (NBO) (FD) Occupancies NBO BD C1−C4 BD C6−C8 LP Cl10 LP Cl10 LP Cl10 -0.3972 -0.4784 -0.1479 -0.0101 -0.0150 -0.4575 -0.5310 -0.1756 -0.0093 -0.0187 1.9424 1.9433 1.9204 1.9755 1.9946 89 82 87 Type ∆fi (NBO) (FMO) ∆fi (NBO) (FD) Occupancies BD C8−Cl10 BD C1−C4 BD C6−C8 -0.0237 0.2945 0.3190 0.0107 0.2549 0.2763 0.0210 0.0586 0.1258 TABLE VIII. Bond reactivity indices ∆fi (NBO) for some important NBOs of Fig. 4 (reagent A2 ). Calculations made with Eq. (S14) (supplementary material). NBO 2 16 15 14 Type ∆fi (NBO) (FMO) ∆fi (NBO) (FD) Occupancies NBO BD C1−C2 LP Cl6 LP Cl6 LP Cl6 -0.7451 -0.3236 -0.0122 -0.0272 -0.7451 -0.3763 -0.0366 -0.0011 1.9970 1.9246 1.9754 1.9946 55 51 the same parameters calculated for all the reagents can be seen in ST 2-20. A high fi− (NBO) or fi+ (NBO) index indicates that the NBO is very reactive and that the orbital is very electrophilic (high fi+ (NBO) ) or very nucleophilic (high fi− (NBO) ). For example, NBOs 9, 4, and 23 have the highest values for the fi− (NBO) in Fig. 1(a) and NBOs 2 and 16 in Fig. 2(a). This indicates that they are the most nucleophilic NBOs. On the other hand, NBOs 87 and 82 in Fig. 1(b) and NBO 51 in Fig. 2(b) have the highest fi+ (NBO) values and so they are the most electrophilic. These results are reasonable because the NBOs that show the highest nucleophilic behavior are those that have higher energies and high partial occupancies. On the other hand, the most electrophilic NBOs are those that have less energies and minor partial occupancies. These orbitals often change during the chemical reaction (for example, the rupture and formation of bonds in ST 21 in the supplementary material). Also, Tables V and VI (and ST 2-20 in the supplementary material) show that the FMO Type ∆fi (NBO) (FMO) ∆fi (NBO) (FD) Occupancies BD C2−Cl6 BD C1−C2 -0.0563 0.7894 -0.0489 0.6978 0.0277 0.0736 approximation leads to the same results (qualitatively) as the FDapproximation. They did not display exactly the same values but showed the same tendencies. Often, the fi− (NBO) and fi+ (NBO) values for the FD approximation are lower than the FMO ones (see ST2-20), the Fukui function under FD approximation has negative zones, and under FMO approximation not. This affects the coefficients but qualitatively it does not change the conclusions obtained with both approximations. B. Testing the bond reactivity index ∆ fi (NBO) based on dual descriptor Figs. 3 and 4 show the ∆fi (NBO) indices obtained with Eq. (27) for the A1 and A2 reagents, where the dual descriptor function (∆f (r)) is used instead of the Fukui function. Tables VII and VIII show the ∆fi (NBO) indices for some important NBOs of Figs. 3 and 4 (the ST 41-57 in the supplementary material show the same parameters for the B1 -F2 reagents). Regarding the meaning of the index, a very small value TABLE IX. fk(NBO) parameter (Eq. (15)) for the most important NBOs of Fig. 1 (reagent A1 ). i Atom number 1 2 3 4 5 6 7 8 9 10 NBO 9 NBO 4 NBO 87 NBO 82 Z BD C6−C8 BD C1−C4 BD C6−C8 BD C1−C4 6 1 1 6 1 6 1 6 1 17 0.000 0.000 0.000 0.010 0.000 0.471 0.000 0.513 0.000 0.005 0.487 0.000 0.000 0.501 0.000 0.011 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.007 0.000 0.516 0.000 0.473 0.000 0.003 0.481 0.000 0.000 0.511 0.000 0.008 0.000 0.000 0.000 0.000 Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 150.214.4.235 On: Mon, 05 Dec 2016 07:54:38 194105-8 Jesús Sánchez-Márquez J. Chem. Phys. 145, 194105 (2016) TABLE X. fk(NBO) parameters (Eq. (15)) for the most important NBOs of i Fig. 2 (reagent A2 ). NBO 2 NBO 51 Atom number Z BD C1−C2 BD C1−C2 1 2 3 4 5 6 6 6 1 1 1 17 0.479 0.517 0.000 0.000 0.000 0.004 0.501 0.495 0.000 0.000 0.000 0.003 (negative) indicates a nucleophilic behavior and a very large index shows an electrophilic behavior. The results obtained for bond reactivity index based on the dual descriptor the ∆fi (NBO) are consistent with the bond reactivity indices based on Fukui function ones (fi− (NBO) and fi+ (NBO) in Tables V and VI). Both indices determine the most reactive orbitals and both methodologies predict the expected reactivity for the D-A reaction of ST 21 (supplementary material). An excellent source for the comparison of previous results is Refs. 32, 48, and 49. As an example, ST 21 in the supplementary material shows some points of the intrinsic reaction coordinate (IRC) for the D-A reaction: A1 + A2 . The NBOs included in ST 21 (NBOs 4 and 9 from A1 and NBO 2 from A2 ) obtained the highest values for the fi− (NBO) , fi+ (NBO) , and ∆fi (NBO) indices (Tables V–VIII). The orbitals related to the new bonds are also included. It is seen that NBOs 4 and 9 from A1 and NBO 2 from A2 disappear in step 39/81 (transition state-1), and then new NBOs appear (to the right in ST 21). Of course, these new NBOs are related to the formation of the new bonds. C. Theoretical regioselectivity The indices fi− (NBO) and fi+ (NBO) calculated in Sec. V A (Figs. 1 and 2) showed that the higher indices (for the A1 reagent) belong to NBOs 4 and 9 (see fi− (NBO) ) and NBOs 82 and 89 (see fi+ (NBO) ), and we will accept that these are the most FIG. 5. Schematic representations of the main NBOs studied. In this case, the arrows represent theoretical polarization of the orbitals due to EWGs or EDGs (conclusions based on the parameters fk(NBO) ). i Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 150.214.4.235 On: Mon, 05 Dec 2016 07:54:38 194105-9 Jesús Sánchez-Márquez reactive orbitals (more nucleophilic and electrophilic). Regarding A2 , the higher fi− (NBO) corresponds to NBO 2 and the higher fi+ (NBO) to NBO 51. The parameter fk(NBO) Eq. (15) i will help to study the regioselectivity of the D-A reaction. This estimates how a NBO is “polarized.” Tables IX and X for the most reactive NBOs of A1 show calculated fk(NBO) i and A2 (main NBOs in Figs. 1 and 2). Tables ST 22-40, in the supplementary material, show calculated parameters for all the reagents of this study. The fk(NBO) parameters depend i on the electron withdrawing and donating groups. For example, in Table IX (A1 reagent), the fk(NBO) parameters for C6 i and C8 indicate that NBO 9 (BD C −C 6 8 ) is displaced to (NBO) C8 f6(NBO) < f and NBO 87 (BD C 6 −C8 ) is displaced 89 9 (NBO) (NBO) to C6 f8 87 < f6 87 . The parameters for C1 and C4 indicate that NBO 4 (BD C1 −C4 ) and NBO 82 are displaced to C4 (f1(NBO) < f4(NBO) and 4 4 (NBO) (NBO) f1 82 < f4 82 ). Regarding A2 , Table X shows that NBO 2 (BD C1 −C2 ) is displaced to C2 and NBO 51 (BD C1 −C2 ) is displaced to C1 . The general conclusions for the four samples (Tables IX and X; ST 22-40 in the supplementary material) are summarized in Fig. 5. Taking into account Fig. 5, we can imagine the charge transfer between the NBOs of A1 and A2, and as a consequence, the theoretical regioselectivity of Fig. 6 can be obtained. The general information obtained from all the samples (ST 2-40 in the supplementary material) is shown in Fig. 6 (and SF 2-5 in the supplementary material), where the arrows represent the charge transfer between a very nucleophilic orbital and a very electrophilic one. The theoretical regioselectivity showed in Fig. 6 and SF 2-5 (donation and acceptance of electron pairs) is as expected for this wellknown reaction and allows us to predict most of the products of the reactions.46,47 Notice that the systems E2 and F2 have a LUMO with positive orbital energies. A new electron added to the molecule might be located in most diffuse orbital of the J. Chem. Phys. 145, 194105 (2016) system giving in these two cases meaningless results for the prediction of a nucleophilic attack. D. Comparison of atomic and bond reactivity indices Under the FMO approximation, Eq. (19) can be written as φ(HOMO) (r)2 ≈ X i 2 fi− (NBO) · φ(NBO) (r) . i (34) 2 2 Condensing the φHOMO (r) and φi(NBO) (r) functions by means of Eqs. (10) and (15), we obtain Eq. (35), X ) fi− (NBO) · fk(NBO . (35) fk− ≈ i i Eq. (35) allows us to relate the atomic reactivity indices fk− (Table XI) and the bond reactivity indices fi− (NBO) . This leads us to the useful conclusion that the atomic reactivity index is the sum of several contributions: the bond reactivity indices of each orbital of the molecule. We think that the fk(NBO) coeffii − (NBO) cients (Table XII) could determine what portion of the fi index belongs to each atom. Also, Eq. (35) is a direct connection between atomic reactivity and bond reactivity indices and can be useful for comparing both. Table XI shows that the atomic reactivity index of the chlorine atom is similar to the carbons (C1 , C4 , C6 , and C8 ). This seems incoherent because chlorine does not react (at least not like the carbons). However, the bond reactivity index of NBO 4 (or NBO 9) is in the order of 1:2 regarding the chlorine NBO 23. Eq. (35) allows us to analyze why different conclusions were obtained from the two types of indices: the double-bonds (NBO 9: BD C6 ==C8 and NBO 4: BD C1 ==C4 in Table V) have a bond reactivity index bigger than the lone-pair of the chlorine (NBO 23). The double bond index is distributed between the two carbons (equally shown in Table XII). On the divided, as the parameter fk(NBO) i other hand, NBO 23 (LP) “only belongs” to the chlorine and FIG. 6. Theoretical donation and acceptance of electron pairs at the transition state (based on the parameters fi− (NBO) , fi+ (NBO) , and fk(NBO) ). Reagents: diene i bearing an EWG at C1 and dienophile bearing an EWG. Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 150.214.4.235 On: Mon, 05 Dec 2016 07:54:38 194105-10 Jesús Sánchez-Márquez J. Chem. Phys. 145, 194105 (2016) FIG. 7. Relationship between atomic reactivity indices (fk− ) for some dienes (A1 , B1 , K1 , and H1 ) in some particular reactions. only contributes to the chlorine index, and this justifies why the atomic reactivity index of the chlorine is bigger than expected. Table XI shows that the fk− indices obtained with Eq. (9) are very different to the ones obtained with Eq. (35), and they even have a different tendency. The atoms with the highest reactivity indices must form the new bonds. For example, in the case of the reagent A1, the double bonds (NBOs) C1==C4 and C6==C8 are the most nucleophile. C4 and C8 are the most nucleophiles atoms (fk− (C1 ) < fk− (C4 ) and fk− (C6 ) < fk− (C8 ) according to Eq. (35)). This could show that atom C8 attacks the dienophile and the C4 the atom C6. Then we obtain the bonds: C8-dienophile (simple) and C4==C6 (double). To close the ring, the most nucleophile atom of the dienophile (that is bonded to the EWG) attacks to the C1 and the bond C1dienophile (simple) is assembled. This interpretation of the regioselectivity based on the indices fk− (Eq. (35)) (and we have tried to summarize in Fig. 7) is coherent with the experimental results. The B1 (ST 58), H1 (ST 63), K1 (ST 67), and many other cases (ST 58-69 of the supplementary material) show similar data behaviour. Broadly speaking, the new index calculated with Eq. (35) shows better results than the other studied indices (Eq. (9) and populations of the NBO analysis). Probably, electron withdrawing and donating groups polarize electronic density and this changes the orbitals (canonical and NBOs), but does not change all the occupied orbitals in the same way. The “standard methodologies” (Eqs. (6)–(8)) use the information of only one frontier orbital TABLE XI. Condensed Fukui function fk− calculated with Eq. (9) (frontier molecular orbital approximation), finite difference (FD) approximation, and atomic population from a NBO analysis, and Eq. (35), for the reagent A1 . Atom 1 2 3 4 5 6 7 8 9 10 fk− fk− fk− Z FMO Eq. (10) FD (NBO analysis) FMO Eq. (35) 6 1 1 6 1 6 1 6 1 17 0.236 0.001 0.001 0.105 0.005 0.177 0.002 0.236 0.001 0.234 0.234 0.025 0.044 0.055 0.055 0.116 0.050 0.130 0.036 0.255 0.185 0.000 0.000 0.194 0.000 0.213 0.000 0.231 0.000 0.186 Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 150.214.4.235 On: Mon, 05 Dec 2016 07:54:38 194105-11 Jesús Sánchez-Márquez J. Chem. Phys. 145, 194105 (2016) TABLE XII. fk(NBO) parameter in Eq. (15) for the most nucleophile NBOs of i reagent A1 . NBO 21 NBO 22 NBO 23 Atom number 1 2 3 4 5 6 7 8 9 10 Z 6 1 1 6 1 6 1 6 1 17 NBO 9 NBO 4 LP Cl10 LP Cl10 LP Cl10 BD C6−C8 BD C1−C4 0.000 0.000 0.000 0.000 0.000 0.003 0.000 0.040 0.001 0.955 0.000 0.000 0.000 0.000 0.000 0.002 0.000 0.023 0.001 0.974 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.014 0.000 0.985 0.000 0.000 0.000 0.010 0.000 0.471 0.000 0.513 0.000 0.005 0.487 0.000 0.000 0.501 0.000 0.011 0.000 0.000 0.000 0.000 (HOMO or LUMO) and there is no guarantee that this orbital was polarized in the same way as the other orbitals (especially, if it has a different symmetry). The bond reactivity methodology (with NBOs) uses the information of all the orbitals and they are also localized orbitals and, consequently, more independent (not like the canonical ones). A perturbation in a localized area of the molecule only affects some bond orbitals or only one (canonical orbital behaviour is different). VI. CONCLUSIONS A new descriptor has been defined: “bond reactivity index.” Considering that a chemical reaction is determined by the breaking and creation of bonds, the new index has the advantage that it is related with bonds instead of atoms. These bond reactivity indices based on the Fukui function and the dual descriptor function have been calculated and the results obtained from the two kinds of indices have been equated. The new indices have been tested with a set of representative molecules and the Diels-Alder reaction type, and we have found a good description of the reactivity. Also, the FD and FMO approximations have been compared and both showed the same tendencies. A new atomic reactivity index has been defined and it provided the best results for the sample of molecules studied. Also, it provided a relationship that we used to compare the atomic reactivity and bond reactivity indices, obtaining useful information about the relationship between the two kinds of indices. SUPPLEMENTARY MATERIAL See supplementary material for the complete reactivity index data of the studied reagents. ACKNOWLEDGMENTS Calculations were performed through CICA (Centro Informático Cientı́fico de Andalucı́a). We thank Professor M. Fernández Núñez, whose suggestions were greatly appreciated. 1 P. Pérez, L. R. Domingo, M. Duque-Noreña, and E. Chamorro, J. Mol. Struct.: THEOCHEM 895, 86 (2009). 2 T. Mineva and N. Russo, J. Mol. Struct.: THEOCHEM 943, 71 (2010). 3 L. R. Domingo, E. Chamorro, and P. Pérez, Eur. J. Org. Chem. 2009, 3036 (2009). 4 P. Bultinck, S. Fias, C. V. Alsenoy, P. W. Ayers, and R. Carbó-Dorca, J. Chem. Phys. 127, 034102 (2007). 5 J. Sánchez-Márquez, J. Mol. Model. 21, 82 (2015). 6 E. Relan and A. E. Reed, Chem. Rev. 88, 899 (1988). 7 E. D. Glendening, R. Landis, and F. Weinhold, WIREs Compt. Mol. Sci. 2, 1 (2012). 8 N. R. Draper and H. Smith, “Applied regression analysis,” Wiley Series in Probability and Statisticsm, 3rd ed. (John Wiley & Sons, 1998), ISBN: 0-471-17082-8. 9 P. Bultinck, S. Van Damme, and A. Cedillo, J. Comput. Chem. 34, 2421 (2013). 10 R. G. Parr, L. V. Szentpaly, and S. B. Liu, J. Am. Chem. Soc. 121, 1922 (1999). 11 R. Parr and W. Yang, Density-Functional Theory of Atoms and Molecules (Oxford University Press, 1989). 12 R. S. Mullikulliken, J. Chem. Phys. 2, 782 (1934). 13 R. P. Iczkowski and J. L. Margrave, J. Am. Chem. Soc. 83, 3547 (1961). 14 K. D. Sen and C. K. Jørgensen, Electronegativity, Structure, and Bonding (Springer, Berlin, 1987). 15 R. G. Pearson, Chemical Hardness: Application from Molecules to Solids (Wiley-VCH, Weinheim, 1997). 16 P. Geerlings, F. De Proft, and W. Langenaeker, Chem. Rev. 103, 1793 (2003). 17 S. Liu, Chemical Reactivity Theory a Density Functional View (CRC Press, 2009), Chap. 13. 18 L. R. Domingo, J. A. Sáez, and P. Pérez, Chem. Phys. Lett. 438, 341 (2007). 19 R. G. Parr and W. Yang, J. Am. Chem. Soc. 106, 4049 (1984). 20 W. Yang and R. G. Parr, Proc. Natl. Acad. Sci. U. S. A. 82, 6723 (1985). 21 W. Yang and W. J. Mortier, J. Am. Chem. Soc. 108, 5708 (1986). 22 F. Zielinski, V. Tognetti, and L. Joubert, Chem. Phys. Lett. 527, 67 (2012). 23 R. R. Contreras, P. Fuentealba, M. Galván, and P. Pérez, Chem. Phys. Lett. 304, 405 (1999). 24 E. Chamorro and P. Pérez, J. Chem. Phys. 123, 114107 (2005). 25 F. Mendizabal, D. Donoso, and D. Burgos, Chem. Phys. Lett. 514, 374 (2011). 26 P. Pérez and R. Contreras, Chem. Phys. Lett. 293, 239 (1998). 27 D. Burgos, C. Olea-Azar, and F. Mendizabal, J. Mol. Model. 18, 2021 (2012). 28 J. Sánchez-Márquez, D. Zorrilla, A. Sánchez-Coronilla, D. M. de los Santos, J. Navas, C. Fernández-Lorenzo, and R. Alcántara, J. Martı́n-Calleja, J. Mol. Model. 20, 2492 (2014). 29 P. C. Hausen, V. Pereyra, and G. de la Scherer, Least Squares Data Fitting with Applications (The Johns Hopkins University Press, 2013), ISBN 13: 978-1-4214-0786-9. 30 C. Morell, A. Grand, and A. Toro-Labbé, J. Phys. Chem. A 109, 205 (2005). 31 V. Tognetti, C. Morell, P. W. Ayers, L. Joubert, and H. Chermetteb, Phys. Chem. Chem. Phys. 15, 14465 (2013). 32 V. Tognetti, C. Morell, and L. Joubert, J. Comput. Chem. 36, 649 (2015). 33 A. D. Becke, J. Chem. Phys. 98, 5648 (1993). 34 M. J. Frisch, J. A. Pople, and J. S. Binkley, J. Chem. Phys. 80, 3265 (1984). 35 M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J. A. Montgomery, Jr., J. E. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador, J. J. Dannenberg, S. Dapprich, A. D. Daniels, O. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski, and D. J. Fox, gaussian 09, Revision A.02, Gaussian, Inc., Wallingford CT, 2009. 36 A. E. Reed, R. B. Weinstock, and F. Weinhold, J. Chem. Phys. 83, 735 (1985). 37 W. Carruthers, Modern Methods of Organic Synthesis (Cambridge University Press, Cambridge, 1978). 38 W. Carruthers, Cycloaddition Reactions in Organic Synthesis (Pergamon Press, Oxford, 1990). Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 150.214.4.235 On: Mon, 05 Dec 2016 07:54:38 194105-12 39 I. Jesús Sánchez-Márquez Fleming, Pericyclic Reaction (Oxford University Press, Oxford, 1999). L. Holmes, Organic Reactions (John Wiley & Sons, New York, 1948). 41 W. Oppolzer, “Intermolecular Diels-Alder reactions,” in Comprehensive Organic Synthesis (Pergamon Press, Oxford, 1991), Vol. 5, p. 315. 42 A. Kumar, Chem. Rev. 101, 1 (2001). 43 S. Damoun, G. Van de Woude, F. Méndez, and P. Geerlings, J. Phys. Chem. A 101, 886 (1997). 44 A. K. Chandra and M. T. Nguyen, J. Phys. Chem. A 102, 6181 (1998). 45 A. Ponti, J. Phys. Chem. A 104, 8843 (2000). 40 H. J. Chem. Phys. 145, 194105 (2016) 46 F. A. Carey and R. J. Sundberg, Advanced Organic Chemistry. Part A: Structure and Mechanisms (Springer Science + Busisness Media, 2007), Vol. 9. 47 F. A. Carey and R. Sundberg, Advanced Organic Chemistry, 3rd ed. (Springer, 1990). 48 C. Morell, P. W. Ayers, A. Grand, S. Gutiérrez-Oliva, and A. Toro-Labbé, Phys. Chem. Chem. Phys. 10, 7239 (2008). 49 P. Geerlings, P. W. Ayers, A. Toro-Labbé, P. K. Chattaraj, and F. De Proft, Acc. Chem. Res. 45, 683 (2012). Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 150.214.4.235 On: Mon, 05 Dec 2016 07:54:38
© Copyright 2025 Paperzz