Skip to main content

ORIGINAL RESEARCH article

Front. Chem., 02 February 2022
Sec. Chemical Physics and Physical Chemistry

How the Nature of an Alpha-Nucleophile Determines a Brønsted Type-Plot and Its Reaction Pathways. An Experimental Study

  • 1Centro de Química Médica, Instituto de Ciencias e Innovación en Medicina, Clínica Alemana Universidad del Desarrollo, Santiago, Chile
  • 2Facultad de Química y de Farmacia, Pontificia Universidad Católica de Chile, Santiago, Chile

The reactions between 2-chloro-5-nitro pyrimidine with a serie of α-nucleophile derivatives were kinetically evaluated. The kinetic study was carried out in aqueous media and the data shown an unusual split on the Brønsted type-plot, opening a controversial discussion based on reactivities and possible reaction pathways. These split Brønsted type-plots are discussed over the hypothetical transition state (TS) structures associated to concerted or stepwise mechanisms with emphasis on hydrogen bond interactions between electrophile/nucleophile pair able to determine the reactivities and the plausible reaction routes.

Introduction

The alpha effect accounts for the increased nucleophilic strength due to the presence of an adjacent atom to the nucleophilic center with a lone pair of electrons (Jencks and Carriuolo, 1960a, 1960b; Edwards and Pearson, 1962; Dixon and Bruice, 1972; Buncel and Um, 2004; Kirby et al., 2005; Ren and Yamataka, 2007; 2009; Ormazábal-Toledo et al., 2013b; Kool et al., 2014; Kölmel and Kool, 2017) The nucleophilic strength has been frequently related with the basicity of the nucleophile. However, sometimes the nucleophilicity is greater than the basicity (Anderson and Jencks, 1960) The nucleophilicity concept is associated to electron-rich species (nucleophiles), at the same way, the electrophilicity to electron-deficient species (electrophile) (Ingold, 1929, 1933, 1934) Both concepts are based on electron theory of Lewis (Lewis, 1923) and the general acid-base theory of Brönsted and Lowry (Brönsted, 1923; Lowry, 1923) Then, nucleophilicity/electrophilicity have been used as quantitative scales in order to rationalize the chemical reactivity (Contreras et al., 2003)

The term “α-effect” was used by Edwards and Pearson in order to describe an additional factor relative to the polarizability that influences the nucleophilicity (Edwards and Pearson, 1962) Currently, there are different hypotheses about this effect, such as: 1) increased polarization of the nucleophiles; 2) stabilization of the Transition State (TS) structures along the of the Potential Energy Surface (PES) by the lone pair at α position; 3) relative stability of the reaction products and 4) ground state destabilization due to electron-electron repulsion (Anderson and Jencks, 1960; Edwards and Pearson, 1962; Dixon and Bruice, 1972; Bell et al., 1974; Fountain et al., 2003; Um et al., 2006; Gallardo-Fuentes et al., 2014) Hudson et al showed that the magnitude of the α-effect will increase with larger βnuc values from Brønsted type-plots (Filippini and Hudson, 1972; Buncel et al., 1993; Fountain et al., 2003) Furthermore, the α-effect is highly modulated by the solvent, but the effect of solvation on the ground state could not explain the changes in the α-effect at higher concentrations of DMSO (Um et al., 1998, 2006) Studies in gas phase have shown that an enhanced α-effect is observed with: 1) high electron density at the α-atom and high electrophilicity values of the electrophile and 2) electronegative α-atom adjacent to the nucleophilic center. However, α-electron withdrawing group diminishes the α-effect (Evanseck et al., 1987; Ren and Yamataka, 2006, 2007; Nigst et al., 2012) Finally, TS structures analysis have shown that there is no difference between nucleophiles with and without α-effect (Ren and Yamataka, 2007)

Therefore, it is possible that the α-effect could be related with several factors, and more studies are needed to provide a detailed description about how this significant effect operates. For better understanding the α-effect, in the present work we studied the magnitude of the α-effect of the reacting pair (electrophile/nucleophile) evaluating the nucleophilic rate coefficients of a nucleophilic aromatic substitution (SNAr) reaction in aqueous media (Cho et al., 2014) The postulated mechanism for a SNAr reaction involves a nucleophilic addition followed by elimination of a leaving group (LG) and it requires the presence of at least one strong electron-withdrawing (Bunnett and Morath, 1955; Liebman et al., 1996) substituent in the ring of the electrophile to stabilize the intermediate, called Meisenheimer Complex (MC) and good LG (Bunnett and Zahler, 1951; Banjoko and Babatunde, 2004; Crampton et al., 2004, 2007; Um et al., 2007; Terrier, 2013; Ormazábal-Toledo et al., 2013b; Gallardo-Fuentes et al., 2014; Gazitúa et al., 2014; Contreras et al., 2015; Calfumán et al., 2017; 2018; Sánchez et al., 2018b) The first step of the reaction mechanism corresponds to the formation of a MC. In a second step, the LG detaches after an intramolecular proton transfer (RLPT) from the nucleophile(Bernasconi and De Rossi, 1976; Ma̧kosza, 1993; Bernasconi et al., 2004; Nudelman, 2009; Um et al., 2012; Ormazabal-Toledo et al., 2013; Ormazábal-Toledo et al., 2013a; Swager and Wang, 2017) Scheme 1 shows the general reaction mechanism for a SNAr reaction. However, more recently, a concerted mechanism has been postulated for this type of reactions. In many cases, the nucleophilic attack on the ipso carbon at the aromatic ring occurs concertedly with the LG departure within a single stepwise pathway without a MC formation (Terrier, 2013; Neumann et al., 2016; Calfumán et al., 2017; Neumann and Ritter, 2017; Stenlid and Brinck, 2017; Kwan et al., 2018; Campodónico et al., 2020) The literature summarizes the mechanistic trends based on the chemical nature of substrates and nucleophiles(Ormazábal-Toledo et al., 2013b; Gazitúa et al., 2014; Alarcón-Espósito et al., 2015, 2016, 2017; 2018; Sánchez et al., 2018a; Campodónico et al., 2020) However, few articles highlight the stabilization of the species along the PES based on hydrogen bond (HB) interactions of the reacting pair (Newington et al., 2007; Ormazábal-Toledo et al., 2013a, 2013b; Gallardo-Fuentes et al., 2014; Calfumán et al., 2017; Sánchez et al., 2018b)

SCHEME 1
www.frontiersin.org

SCHEME 1. General reaction mechanism for a SNAr with a hypothetical protonated nucleophile. LG corresponds to the Leaving Group and EWG corresponds to electron withdrawing groups.

In this work, we studied the reaction of 2-chloro-5-nitro pyrimidine (electrophile) with the family of α-nucleophiles depicted in Table 1 (see bottom in Results and Discussion) in aqueous media. Scheme 2 describes the SNAr reaction between 2-chloro-5-nitro pyrimidine and a hypothetical alpha-nucleophile. The main goal was to determine the α-effect on the studied reaction considering the kinetic results and the analysis of the Brønsted type-plot in addition to chemical structures analysis of the reacting pairs.

TABLE 1
www.frontiersin.org

TABLE 1. Summary of nucleophiles and their pKa values in water and kN values for the nucleophile series with 2-chloro-5-nitro pyrimidine.

SCHEME 2
www.frontiersin.org

SCHEME 2. General reaction mechanism for a SNAr between 2-chloro-5-nitro pyrimidine with a hypothetical protonated nucleophile.

A Brønsted plot corresponds to a free energy relationship that correlates the logarithm of the nucleophilic rate coefficients (kN) and the pKa values of the nucleophiles from Brønsted Equation:

logkN=βnuc pKa+log G (1)

where G is a constant that depends of the solvent and temperature and βnuc corresponds to the development of charge between the reaction sites of the nucleophile/electrophile pair, respectively, along to the PES. (Brönsted and Pedersen, 1924) Therefore, βnuc gives information about the TS structure related to the rate determining step (RDS) in the reaction mechanism. (Buncel et al., 1993)

Materials and Methods

Reactants

2-Chloro-5-nitro pyrimidine and all the nucleophiles were of the highest quality available commercial products by Sigma Aldrich and Merck. The certificate of analysis guarantees purity ≥99%.

Kinetic Measurements

The kinetics were carried out spectrophotometrically by means of a diode array spectrophotometer in aqueous media, monitoring the appearance of 2,4-dinitrophenoxide anion at 360 nm. The experimental conditions were 25.0 ± 0.1°C, ionic strength 0.2 M (KCl), at three different pH values maintained by partial protonation of the nucleophiles. All the reactions were studied under excess of the nucleophile at least 10 times greater than the substrate concentration (Um et al., 2007, 2012) in order to achieve pseudo-first-order kinetic conditions. The reactions were started by injection of a substrate stock solution 0.1 M in acetonitrile (10 μl) into the amine solution (2.5 ml in the spectrophotometric cell) reaching a concentration of 0.0004 M in the cell. The formation of colored amino-substituted nitropyrimidine compounds were monitored by UV–vis spectroscopy. In all runs, the pseudo-first-order rate constant (kobs) was found for all the reactions. The kobs were determined by means of the spectrophotometer kinetic software for first order reactions at the wavelength corresponding to the kinetic products. Note that, in aqueous media each pH values correspond to: pH = pKa and 0.3 units up and down in order to analyze the possibility of acid and/or basic catalysis by the reaction media. On the other hand, a Brønsted type-plot requires a broad range of pKa values for the nucleophiles. For this reason, in this study was used a family of nucleophiles with similar chemical features. Then, the relationships between kobs vs [Nu] (nucleophile concentration) should be straight lines or straight lines with smooth deviations, which will discard a catalysis processes by the media. See more details in Supplementary Figures S1-S6 and Supplementary Tables S1-S18, respectively in Supplementary Material (SM). This kinetic methodology was taken from previous kinetic studies cited in literature and previous works performed by our group (Castro et al., 1999; Castro et al., 2007; Um et al., 2007; Ormazábal-Toledo et al., 2013, 2013a, 2013b; Gallardo-Fuentes et al., 2014; Gazitúa et al., 2014; Alarcón-Espósito et al., 2015, 2016,2017; Calfuman et al., 2017; 2018; Sánchez et al., 2018a, 2018b; Campodonico et al., 2020)

Product Analysis

In the studied reactions, the increase of a band centred in the range of 330—550 nm was observed; attributed to the corresponding reaction products for all nucleophile series studied.

Synthesis of Products

5-Nitro-N-phenylpyrimidin-2-amine

To a solution of 2-chloro-5-nitropyrimidine (40 mg, 0.25 mmol) in CH3CN (1.0 ml), was added aniline (23.3 mg, 0.25 mmol). The reaction mixture was stirred for 4 h at room temperature, the solvent was removed under vacuum to give a yellow solid which was recrystallized from ethanol (35 mg, 65%), mp 201.5–202.5°C (Lit (Von Bebenburg and Thiele, 1970) 202–203°C). 1H-NMR (400 MHz, DMSO-d6) d: 7.13 (t, J = 7.5 Hz, 1H), 7.36 (t, J = 8.0 Hz, 2H), 7.76 (d, J = 8.0 Hz, 2H), 9.22 (s, 2H), 10.84 (s, 1H); 13C-NMR (100 MHz, DMSO-d6) d: 126.0, 129.2, 133.9, 140.3, 143.6, 160.3, 166.0.

2-Hydrazinyl-5-Nitropyrimidine

Using the above procedure, from 2-chloro-5-nitropyrimidine (40 mg, 0.25 mmol) and hydrazine (8.0 mg, 0.25 mmol), was obtained a yellow solid (30 mg, 77%), mp 170–172°C (Lit (Caton and McOmie, 1968) 168–169°C). 1H-NMR (400 MHz, DMSO-d6) d: 9.13 (s, 1 H), 9.20 (s, 1 H), 10.84 (s, 1 H); 13C-NMR (100 MHz, DMSO-d6) d: 136.3, 155.9, 164.3.

Results and Discussion

In the experimental conditions used only one product formation was spectrophotometrically observed for all the reactions studied. Therefore, the possibility of nucleophilic attack at the unsubstituted ring positions is discarded (Um et al., 2007) This fact was confirmed by synthesis and study of the reaction product (see Experimental Section and SM), discarding the possibility of nucleophilic attack at the unsubstituted positions on the aromatic ring (4 and 6, positions).

The values of kobs for all the reactions are in accordance with Eq. 2 where k0 and kN are the rate coefficients for hydrolysis and aminolysis, respectively. Then, the kobs values were obtained at different concentrations of the nucleophile in aqueous media. The kobs values were plotted vs [Nu] in order to obtain kN values from Eq. 2:

kobs=  k0 + kN[Nu](2)

The kobs  for the reactions can be expressed as Eq. 3(Terrier, 2013; Contreras et al., 2015) and k1, k2 and k3 are the micro-constants associated to the reaction mechanism of an SNAr reaction (see Scheme 1 and Scheme 2) and obtained applying the steady-state approximation to the SNAr mechanisms:

kobs=(k1k2[N] + k1k3[Nu]2)(k1 + k2 + k3[Nu])(3)

Linear plots of kobs  vs free nucleophile concentration ( [Nu] F) that pass through the origin, suggest that the contribution of the solvent to the kobs values is negligible and the reactions occurs via a non-catalyzed route (k2 route in Scheme 1).(Um et al., 2007; Gazitúa et al., 2014; Sánchez et al., 2018b) Thus, kobs  can be expressed as Eq. 4, where kN is determined from the slope of the linear plots, where k1 + k2 >>> k3 [Nu].

kobs = kN[Nu], where kN = k1k2k1 + k2(4)

The values of kN and pKa are summarized in Table 1 (kinetic details are in Experimental Section and SM). In order to have a reasonable set of nucleophiles of varying basicity (broad range of pKa values) and nucleophilicity, pKa data were taken from the literature (Kirby et al., 2008) The kN and pKa values from Table 1 were statistically corrected with p and q parameters, where q is the number of equivalent basic sites on the free nucleophile, and p is the number of equivalent dissociable protons on the conjugate acid of the nucleophile (Bell, 1973) The values accompanying kN in Table 1 correspond to the error associated to the slope to obtain these kinetic coefficient values.

A preliminary inspection of Table 1 reveals that the general trend in reactivity is: N-methyl hydroxylamine > hydrazine > N,N-dimethyl hydroxylamine > N,O-dimethyl hydroxylamine > hydroxylamine > methoxylamine. Note that, this trend is not in agreement with the pKa values of the α-nucleophiles. These α-nucleophiles that have a lone pair vicinal to the attacking nitrogen atom, should display an enhanced nucleophilicity towards 2-chloro-5-nitro pyrimidine. However, the kinetic data showed that the α-effect in this case is not high. This fact suggests that the solvent has a significant effect over the reaction (Buncel and Um, 2004) Note that, water is a molecule with high capacity to establish HB donor/acceptor, then water molecules could be decreasing the nucleophilicity of these α-nucleophiles.

Figure 1 shows the statistically corrected Brønsted-type plot for the studied reactions, and the nucleophile serie do not follow the same trend. Unusually, the Brønsted-type plot is split in two trends, but three points in each one is not enough to establish a correlation and to establish the rate-determining step (RDS) of the reaction mechanism. However, in a first approach a split Brønsted-type plot would suggest that: 1) the studied nucleophile serie have TS structurally different and they should be associated to RDS of the reaction mechanism and 2) the reactivity of the nucleophiles is associated to its chemical structure and steric hindrance close to the nucleophilic center.

FIGURE 1
www.frontiersin.org

FIGURE 1. Brønsted -type plots (statistically corrected) obtained for the reactions of 2-choro-5-nitro pyrimidine with alpha nucleophile series in aqueous solution, at 25.0°C and ionic strength of 0.2 M in KCl. In increasing order of pKa: empty circles correspond to N,O-dimethyl hydroxylamine, N,N-dimethyl hydroxylamine and N-methyl hydroxylamine compounds; and full circles correspond to: methoxylamine, hydroxylamine and hydrazine compounds.

Then, from Figure 1 is observed an increased order in reactivity for the nucleophiles in both trends in agreement with theirs pKa values. On the other hand, the chemical structure analysis shows the following:

1) The first trend in nucleophilicity is denoted by full circles in Figure 1 that shows the reactivities of hydrazine > hydroxylamine > methoxylamine which agrees with their kN  values. Note that, hydroxylamine is 11.5 times more reactive than methoxylamine and hydrazine 158 times more reactive than methoxylamine. Considering hydrazine as reference compound the influence of the substituent on the nucleophilic reactivity was analyzed. Replacement of one −NH2 group in hydrazine by a −OH group reduces the nucleophilicity, and a similar effect is observed replacing one −NH2 group in hydrazine by a −MeO group. This trend suggests for hydroxylamine and methoxylamine that the oxygen atom adjacent to the nucleophilic center diminishes the reactivity and that the presence of a −CH3 group in methoxylamine diminishes HB ability of the nucleophile. A previous report of the reaction 2-chloro-5-nitro pyrimidine with benzohydrazine derivatives demonstrated that the intramolecular HB enhance the nucleophilicity of these α-nucleophiles (Gallardo-Fuentes et al., 2014) Note that, this HB will be formed by hydroxylamine and hydrazine, respectively toward the substrate (see Scheme 3 below). However, methoxylamine does not have the possibility to establish this HB. This specific interaction would be in the TS structure providing information to explain the kinetic behavior of this trend (see Table 1). The synergy of both HB interactions (oriented to electrophilic centre and LG, respectively) would indicate a concerted route. In agreement with the experimental results, the general-base catalyzed mechanism denoted by k3 [Nu] in Scheme 1 and Scheme 2 is excluded. Then, the possibility of a stepwise mechanism is still open (k2 channel in Scheme 1 and Scheme 2). This HB interaction will promote the electron delocalization on the pyrimidine moiety activating the electrophile and nucleophilicity of the α-nucleophile. Then, the nucleophilicity of the α-nucleophile added to the high nucleofugality of the LG of the heterocyclic ring suggests that the MC intermediate is not stable and the reaction mechanism proceeds through one TS structure and a concerted route is suggested. It is interesting to note that Kwan et al. recently suggested that heterocycles that contain nitrogen atoms and good LG follow a concerted trend (Kwan et al., 2018) Furthermore, Campodonico et al. proposed a concerted mechanism for the reaction of 2-chloro-5-nitro pyrimidine with primary and secondary alicyclic amines (Campodónico et al., 2020) Moreover, Bernasconi et al. postulated that the existence of an intramolecular HB between a hydrogen atom of the nucleophilic centre (amine) and the o-NO2 group of the substrate could explain the reactivity trend (Bernasconi et al., 2004; Ormazábal-Toledo et al., 2013b) In addition, computational reports based on experimental studies emphasize the role of HB on activating the reacting pair (electrophile/nucleophile) and stabilizing the TS (Bunnett and Morath, 1955; Zingaretti et al., 2003; Bernasconi et al., 2004; Gordillo et al., 2007; Alvaro et al., 2011; Ormazábal-Toledo et al., 2013b; Gallardo-Fuentes et al., 2014; Rohrbach et al., 2020)

SCHEME 3
www.frontiersin.org

SCHEME 3. Possible HB interaction between the reacting pair. Structures correspond to hydrazine (A), hydroxylamine (B) and methoxylamine (C) nucleophiles toward 2-chloro-5-nitro pyrimidine, respectively.(Gallardo-Fuentes et al., 2014)

2) The second trend (empty circles in Figure 1) shows the following order of reactivity: N-methyl hydroxylamine > N,N-dimethyl hydroxylamine > N,O-dimethyl hydroxylamine. This trend shows the decreasing effect of methyl groups on the nucleophilic reactivity; N-methylhydroxylamine is 2.3 times more reactive than N,N-dimethylhydroxylamine, which in turns is 4.6 times more reactive than N,O-dimethylhydroxylamine.

The comparison between both trends shown an increase in reactivity for the second trend (see Figure 1). For instance, N,N-dimethyl hydroxylamine and hydroxylamine have similar pKa values, but the first increased its rate coefficient value in 9 times. This fact suggests that the inductive effect of methyl group on these structures play a key role in the reactivity of this trend stabilizing the ammonium cation in the TS structures, enhancing the reactivity of the nucleophiles promoting the nucleophilic attack. But, this stabilizing effect could be diminished by steric hindrance in N,N-dimethyl hydroxylamine. The observed effects that methyl groups increase the nucleophilicity of the substituted nitrogen and decrease the reactivity of the adjacent center was described before by Nigst et al. (Nigst et al., 2012) Furthermore, the HB interaction, is activating the electrophile and nucleophilicity of the α-nucleophile, except in N,O-dimethyl hydroxylamine. Thus, in this second trend the nucleophilicity strength is higher toward the substrate. See Scheme 4.

SCHEME 4
www.frontiersin.org

SCHEME 4. Possible HB interaction between the reacting pair. Structures correspond to N-methyl hydroxylamine (D), N,N-dimethyl hydroxylamine (E) and N,O-dimethyl hydroxylamine (F) nucleophiles toward 2-chloro-5-nitro pyrimidine, respectively. (In analogy to Scheme 2) (Gallardo-Fuentes et al., 2014)

In order to reinforce the hypothesis that stereo-electronic effects on TS stabilization, may activate the electrophile and to improve the nucleophilicity of the nucleophile, a kinetic study of a serie of anilines using the same substrate was performed. With this purpose, the stereo-electronic effect of electron-donors (-NH2, -OMe, -Me) and one electron-acceptor (Me-C=O) groups in the nucleophile, was studied. Table 2 summarize the values of kN and pKa (kinetic details are in Experimental Section and SM). Plots of kobs vs [Nu] shown straight lines in accordance with Eq. 3, thereby indicating that the reaction proceeds through a non-catalyzed mechanism (k2 channel in Scheme 1 and Scheme 2 and Supplementary Figures S7-S12 and Supplementary Tables S19-S26 for more details in SM). The pKa data were taken from the literature in order to have a reasonable set of nucleophiles of varying basicity and nucleophilicity (Castro et al., 1999) The kN and pKa values from Table 2 were statistically corrected with p and q parameters, respectively (Bell, 1973) The values accompanying kN in Table 2 correspond to the error associated to the slope, respectively to obtain these kinetic coefficient values.

TABLE 2
www.frontiersin.org

TABLE 2. Aniline serie and their pKa values in water and kN values for the nucleophile series with 2-chloro-5-nitro pyrimidine.

Figure 2 shows a Brønsted type-plot with a βnuc value of 0.83 for the aniline serie (full squares). This value suggests that the bond formation between the nucleophile (aniline derivatives) and the substrate is fully advanced in the rate-limiting TS. This value agrees with the βnuc value reported for the SNAr reaction between 2,4-dinitrophenylsulfonylchloride with secondary alicyclic (SA) amines in aqueous media, where the LG departure was attributed as the RDS for a non-catalyzed pathway (Gazitúa et al., 2014) This fact, would suggest a stepwise route where the LG departure is the RDS on the reaction mechanism for the aniline serie. Then, the unusual split Brønsted-type plot for the alpha nucleophile above (second trend for empty circles in Figure 1) reinforce the idea that it will be associated to a change on the reaction pathway for the studied nucleophile series toward the substrate; suggesting a stepwise mechanism were the RDS is LG departure (See Figure 2).

FIGURE 2
www.frontiersin.org

FIGURE 2. Brønsted -type plot (statistically corrected) obtained for the reactions of 2-choro-5-nitro pyrimidine with aniline series in aqueous solution, at 25.0°C and ionic strength of 0.2 M in KCl (full square). The empty circles correspond to: N,O-dimethyl hydroxylamine, N,N-dimethyl hydroxylamine and N-methyl hydroxylamine compounds and full circles correspond to: methoxylamine, hydroxylamine and hydrazine compounds, respectively (see Figure 1).

Focusing our analyses over the chemical structure of the aniline serie; the rate coefficients are notably sensitive to the inductive effects of the substituents. Thus, electron-donating p-substituent has a strong effect on the nucleophilicity and the reactivity order for the nucleophiles agrees with theirs pKa values (see Table 2). For instance, 4-phenylenediamine (pKa  = 6.20) has the highest nucleophilic rate coefficient and 3-aminoacetophenone (pKa  = 3.64) the lowest. Therefore, electron-donating substituent plays an important role on the stabilization of the positive charge on the anilinium cation TS structure (see Scheme 5 above). Then, hydrogen-bonding interactions of the media (solvent as acceptor with β parameter) with positive charge on the activated complex of the reaction will stabilize the activated complex better than the reactants; therefore, increasing the β parameter accelerates the reaction rate (Kamlet et al., 1983)

SCHEME 5
www.frontiersin.org

SCHEME 5. A general scheme of anilinium cation TS structure.

Accordingly, heterocyclic substrates that contains nitrogen atoms in its chemical structure assist a favorable nucleophilic attack by high nucleophilic amines, but slow LG departures. On this way, the nature of the reacting pair and the reaction media drastically affects the nucleophilic reaction rates and the RDS on the reaction mechanism (Klopman and Frierson, 1984; Garver et al., 2011; Gazitúa et al., 2018; Um et al., 2018)

Finally, in order to determine the HB effect, it was carried out the kinetic study of phenyl hydrazine (see Supplementary Tables S37-S39 and Supplementary Figure S13 in SM) with the same substrate. Note that, this nucleophile with a potential alpha-effect showed a similar behaviour than aniline derivatives (kN = 1.45 ± 0.006 M−1s−1 and pKa = 5.25 versus kN = 0.99 ± 0.0139 M−1s−1 and pKa = 4.73, respectively) reinforcing the substituent effect exerted over the nucleophilic center. Moreover, is interesting to analyze the nucleophilic rate values for phenyl hydrazine and p-phenylenediamine compounds (kN = 1.45 ± 0.006 M−1s−1 and pKa = 5.25 versus kN = 33.7 ± 0.0610 and pKa = 5.25, respectively) (Brighente and Yunes, 1997) Therefore, in the aniline serie, the fundamental role is played by the inductive effect of the substituents increasing the nucleophilicity and stabilizing the anilinium cation TS structure.

Concluding Remarks

A complete experimental study on an SNAr reaction has been presented. The experimental results shown an unusual broken on the Brönsted type-plot for the alpha nucleophiles studied, suggesting TS structures structurally different given by the reactivities associated to the chemical structure of them: First, an HB interaction is suggested between the α-hydrogen atom of the nucleophile which is oriented toward the nitrogen atom of the pyrimidine moiety. This HB will promote the reactivity of this serie. Then, a second HB oriented towards the LG, added to the chemical features of the reacting pairs, suggest a concerted route. The second family of alpha-nucleophiles showed a key role of the methyl group inductive effect, stabilizing the ammonium cation in the TS structures, and increasing the reactivity of the nucleophiles. Then, a complete kinetic study based on aniline derivatives toward the same electrophile in order to analyze the Brönsted type-plot, observing a high βnuc value. This value suggests that the bond formation between the aniline derivatives and the substrate is fully advanced in the rate-limiting TS and LG departure is the RDS for a non-catalyzed pathway. On the other hand, the stereo-electronic effects on TS stabilization shows that an electron-donating substituent plays an important role on the stabilization of the positive charge on the anilinium cation TS structure accelerating the nucleophilic attack. In summary, the magnitude of the alpha effect depends on the chemical structure of the nucleophiles added to solvent effect, and particularly the possibility to stablish HB interactions between the reacting pair. Then, a detailed experimental study must consider all the factors that are contributing to the reactivity and determining the reaction pathway. An interesting point will be to test these reactions in an aprotic solvents and/or non-conventional solvent such as deep eutectic solvent or an ionic liquid.

Data Availability Statement

The original contributions presented in the study are included in the article/Supplementary Material, further inquiries can be directed to the corresponding author.

Author Contributions

PC design the experiments, performed the kinetic data, analysed results, wrote, discussed and revised the manuscript. RT performed the synthesis and characterization of the reaction products. He discussed and revised the manuscript. CS performed some kinetic data and worked in the manuscript. All the authors have approved the final revised manuscript. PC and behalf of Collaborative Working Group.

Funding

This work was supported by Fondecyt Grant 1150759.

Conflict of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher’s Note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors, and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

Supplementary Material

The Supplementary Material for this article can be found online at: https://www.frontiersin.org/articles/10.3389/fchem.2021.740161/full#supplementary-material

References

Alarcón-Espósito, J., Contreras, R., and Campodónico, P. R. (2017). Iso-solvation Effects in Mixtures of Ionic Liquids on the Kinetics of a Model SNAr Reaction. New J. Chem. 41, 13435–13441. doi:10.1039/C7NJ03246C

CrossRef Full Text | Google Scholar

Alarcón-Espósito, J., Contreras, R., Tapia, R. A., and Campodónico, P. R. (2016). Gutmann's Donor Numbers Correctly Assess the Effect of the Solvent on the Kinetics of SNAr Reactions in Ionic Liquids. Chem. Eur. J. 22, 13347–13351. doi:10.1002/chem.201602237

CrossRef Full Text | Google Scholar

Alarcón-Espósito, J., Tapia, R. A., Contreras, R., and Campodónico, P. R. (2015). Changes in the SNAr Reaction Mechanism Brought about by Preferential Solvation. RSC Adv. 5, 99322–99328. doi:10.1039/C5RA20779G

CrossRef Full Text | Google Scholar

Alvaro, C. E. S., Ayala, A. D., and Nudelman, N. S. (2011). Hydrogen-bonded Nucleophile Effects in ANS: the Reactions of 1-chloro and 1-Fluoro-2,4-Dinitrobenzene with 2-guanidinobenzimidazole, 1-(2-aminoethyl)piperidine andN-(3-Aminopropyl)morpholine in Aprotic Solvents. J. Phys. Org. Chem. 24, 101–109. doi:10.1002/poc.1712

CrossRef Full Text | Google Scholar

Anderson, B. M., and Jencks, W. P. (1960). The Effect of Structure on Reactivity in Semicarbazone Formation1. J. Am. Chem. Soc. 82, 1773–1777. doi:10.1021/ja01492a057

CrossRef Full Text | Google Scholar

Banjoko, O., and Babatunde, I. A. (2004). Rationalization of the Conflicting Effects of Hydrogen Bond Donor Solvent on Nucleophilic Aromatic Substitution Reactions in Non-polar Aprotic Solvent: Reactions of Phenyl 2,4,6-trinitrophenyl Ether with Primary and Secondary Amines in Benzene-Methanol Mixtures. Tetrahedron 60, 4645–4654. doi:10.1016/j.tet.2004.03.079

CrossRef Full Text | Google Scholar

Bell, R. P., Critchlow, J. E., and Page, M. I. (1974). ChemInform Abstract: Ground State and Transition State Effects in the Acylation of Alpha-Chymotrypsin in Organic Solvent-Water Mixtures. Chemischer Informationsdienst 5, no. doi:10.1002/chin.197412124

CrossRef Full Text | Google Scholar

Bell, R. P. (1973). The Proton in Chemistry. 2nd ed. Norwich, Scotland: Chapman & Hall.

Google Scholar

Bernasconi, C. F., Ali, M., Nguyen, K., Ruddat, V., and Rappoport, Z. (2004). Reactions of Substituted (Methylthio)benzylidene Meldrum's Acids with Secondary Alicyclic Amines in Aqueous DMSO. Evidence for Rate-Limiting Proton Transfer. J. Org. Chem. 69, 9248–9254. doi:10.1021/jo040244s

PubMed Abstract | CrossRef Full Text | Google Scholar

Bernasconi, C. F., and De Rossi, R. H. (1976). Influence of the O-nitro Group on Base Catalysis in Nucleophilic Aromatic Substitution. Reactions in Benzene Solution. J. Org. Chem. 41, 44–49. doi:10.1021/jo00863a010

CrossRef Full Text | Google Scholar

Brighente, I. M. C., and Yunes, R. A. (1997). The General Mechanisms of Attack of Nitrogen Nucleophiles on Carbonyl Compounds: Facts that Determine the Change of the Rate-pH Profiles. J. Braz. Chem. Soc. 8, 549–553. doi:10.1590/S0103-50531997000500018

CrossRef Full Text | Google Scholar

Brönsted, J. N. (1923). Einige Bemerkungen über den Begriff der Säuren und Basen. Recl. Trav. Chim. Pays-bas 42, 718–728. doi:10.1002/recl.19230420815

CrossRef Full Text | Google Scholar

Brönsted, J. N., and Pedersen, K. (1924). Die katalytische Zersetzung des Nitramids und ihre physikalisch-chemische Bedeutung. Z. für Phys. Chem. 108U, 185–235. doi:10.1515/zpch-1924-10814

CrossRef Full Text | Google Scholar

Buncel, E., Tarkka, R., and Hoz, S. (1993). The Phenomenology of Differently Constructed Brønsted-type Plots. J. Chem. Soc. Chem. Commun., 1993 109–110. doi:10.1039/C39930000109

CrossRef Full Text | Google Scholar

Buncel, E., and Um, I.-H. (2004). The α-effect and its Modulation by Solvent. Tetrahedron 60, 7801–7825. doi:10.1016/j.tet.2004.05.006

CrossRef Full Text | Google Scholar

Bunnett, J. F., and Morath, R. J. (1955). The Rates of Condensation of Piperidine with 1-Chloro-2,4-Dinitrobenzene in Various Solvents. J. Am. Chem. Soc. 77, 5165. doi:10.1021/ja01624a063

CrossRef Full Text | Google Scholar

Bunnett, J. F., and Zahler, R. E. (1951). Aromatic Nucleophilic Substitution Reactions. Chem. Rev. 49, 273–412. doi:10.1021/cr60153a002

CrossRef Full Text | Google Scholar

Calfumán, K., Gallardo-Fuentes, S., Contreras, R., Tapia, R. A., and Campodónico, P. R. (2017). Mechanism for the SNAr Reaction of Atrazine with Endogenous Thiols: Experimental and Theoretical Study. New J. Chem. 41, 12671–12677. doi:10.1039/C7NJ02708G

CrossRef Full Text | Google Scholar

Campodónico, P. R., Olivares, B., and Tapia, R. A. (2020). Experimental Analyses Emphasize the Stability of the Meisenheimer Complex in a SNAr Reaction toward Trends in Reaction Pathways. Front. Chem. 8, 583. doi:10.3389/fchem.2020.00583

PubMed Abstract | CrossRef Full Text | Google Scholar

Castro, E. A., Cañete, A., Campodónico, P. R., Cepeda, M., Pavez, P., Contreras, R., et al. (2013). Kinetic and Theoretical Study on Nucleofugality in the Phenolysis of 3-nitrophenyl and 4-nitrophenyl 4-cyanophenyl Thionocarbonates. Chem. Phys. Lett. 572, 130–135. doi:10.1016/j.cplett.2013.04.002

CrossRef Full Text | Google Scholar

Castro, E. A. (2007). Kinetics and Mechanisms of Reactions of Thiol, Thiono and Dithio Analogues of Carboxylic Esters with Nucleophiles. An Update. J. Sulfur Chem. 28, 401–429. doi:10.1080/17415990701415718

CrossRef Full Text | Google Scholar

Castro, E. A., Leandro, L., Millán, P., and Santos, J. G. (1999). Kinetics and Mechanism of the Reactions of Anilines with Ethyl S-Aryl Thiocarbonates. J. Org. Chem. 64, 1953–1957. doi:10.1021/jo982063u

PubMed Abstract | CrossRef Full Text | Google Scholar

Caton, M. P. L., and McOmie, J. F. W. (1968). Pyrimidines. Part XVII. Nitration of 5-Acetamido-2-Phenylpyrimidine and the Synthesis of Some 5-nitropyrimidines. J. Chem. Soc. C, 836–838. doi:10.1039/J39680000836

CrossRef Full Text | Google Scholar

Cho, H.-J., Kim, M.-Y., and Um, I.-H. (2014). The α-Effect in SNAr Reaction of Y-Substituted-Phenoxy-2,4-Dinitrobenzenes with Amines: Reaction Mechanism and Origin of the α-Effect. Bull. Korean Chem. Soc. 35, 2448–2452. doi:10.5012/BKCS.2014.35.8.2448

CrossRef Full Text | Google Scholar

Contreras, R., Andres, J., Safont, V. S., Campodonico, P., and Santos, J. G. (2003). A Theoretical Study on the Relationship between Nucleophilicity and Ionization Potentials in Solution Phase. J. Phys. Chem. A. 107, 5588–5593. doi:10.1021/jp0302865

CrossRef Full Text | Google Scholar

Contreras, R., Campodónico, P. R., and Ormazábal-Toledo, R. (2015). “Theoretical and Experimental Methods for the Analysis of Reaction Mechanisms in SNAr Processes,” in Arene Chemistry: Reaction Mechanisms and Methods for Aromatic Compounds. Editor J. Mortier (Wiley & Sons), 175–193. doi:10.1002/9781118754887.ch7

CrossRef Full Text | Google Scholar

Crampton, M. R., Emokpae, T. A., Howard, J. A. K., Isanbor, C., and Mondal, R. (2004). Leaving Group Effects on the Mechanism of Aromatic Nucleophilic Substitution (SNAr) Reactions of Some Phenyl 2,4,6-trinitrophenyl Ethers with Aniline in Acetonitrile. J. Phys. Org. Chem. 17, 65–70. doi:10.1002/poc.690

CrossRef Full Text | Google Scholar

Crampton, M. R., Emokpae, T. A., and Isanbor, C. (20072007). The Effects of Ring Substituents and Leaving Groups on the Kinetics of SNAr Reactions of 1-Halogeno- and 1-Phenoxy-Nitrobenzenes with Aliphatic Amines in Acetonitrile. Eur. J. Org. Chem. 2007, 1378–1383. doi:10.1002/ejoc.200600968

CrossRef Full Text | Google Scholar

Dixon, J. E., and Bruice, T. C. (1972). .alpha. Effect. V. Kinetic and Thermodynamic Nature of the .Alpha. Effect for Amine Nucleophiles. J. Am. Chem. Soc. 94, 2052–2056. doi:10.1021/ja00761a043

CrossRef Full Text | Google Scholar

Edwards, J. O., and Pearson, R. G. (1962). The Factors Determining Nucleophilic Reactivities. J. Am. Chem. Soc. 84, 16–24. doi:10.1021/ja00860a005

CrossRef Full Text | Google Scholar

Evanseck, J. D., Blake, J. F., and Jorgensen, W. L. (1987). Ab Initio study of the SN2 Reactions of Hydroxide and Hydroperoxide with Chloromethane. J. Am. Chem. Soc. 109, 2349–2353. doi:10.1021/ja00242a018

CrossRef Full Text | Google Scholar

Filippini, F., and Hudson, R. F. (1972). A General Treatment of Enhanced Nucleophilic Reactivity. J. Chem. Soc. Chem. Commun., 522–523. doi:10.1039/C39720000522

CrossRef Full Text | Google Scholar

Fountain, K. R., Felkerson, C. J., Driskell, J. D., and Lamp, B. D. (2003). The α-Effect in Methyl Transfers from S-Methyldibenzothiophenium Fluoroborate to Substituted N-Methylbenzohydroxamates. J. Org. Chem. 68, 1810–1814. doi:10.1021/jo0206263

CrossRef Full Text | Google Scholar

Gallardo-Fuentes, S., Tapia, R. A., Contreras, R., and Campodónico, P. R. (2014). Site Activation Effects Promoted by Intramolecular Hydrogen Bond Interactions in SNAr Reactions. RSC Adv. 4, 30638–30643. doi:10.1039/C4RA04725G

CrossRef Full Text | Google Scholar

Garver, J. M., Gronert, S., and Bierbaum, V. M. (2011). Experimental Validation of the α-Effect in the Gas Phase. J. Am. Chem. Soc. 133, 13894–13897. doi:10.1021/ja205741m

CrossRef Full Text | Google Scholar

Gazitúa, M., Tapia, R. A., Contreras, R., and Campodónico, P. R. (2018). Effect of the Nature of the Nucleophile and Solvent on an SNAr Reaction. New J. Chem. 42, 260–264. doi:10.1039/C7NJ03212A

CrossRef Full Text | Google Scholar

Gazitúa, M., Tapia, R. A., Contreras, R., and Campodónico, P. R. (2014). Mechanistic Pathways of Aromatic Nucleophilic Substitution in Conventional Solvents and Ionic Liquids. New J. Chem. 38, 2611–2618. doi:10.1039/C4NJ00130C

CrossRef Full Text | Google Scholar

Gordillo, R., Dudding, T., Anderson, C. D., and Houk, K. N. (2007). Hydrogen Bonding Catalysis Operates by Charge Stabilization in Highly Polar Diels−Alder Reactions. Org. Lett. 9, 501–503. doi:10.1021/ol0629925

PubMed Abstract | CrossRef Full Text | Google Scholar

Ingold, C. K. (1933). 266. Significance of Tautomerism and of the Reactions of Aromatic Compounds in the Electronic Theory of Organic Reactions. J. Chem. Soc., 1120–1127. doi:10.1039/JR9330001120

CrossRef Full Text | Google Scholar

Ingold, C. K. (1934). Principles of an Electronic Theory of Organic Reactions. Chem. Rev. 15, 225–274. doi:10.1021/cr60051a003

CrossRef Full Text | Google Scholar

Ingold, C. K. (1929). The Principles of Aromatic Substitution, from the Standpoint of the Electronic Theory of Valency. Recl. Trav. Chim. Pays-bas 48, 797–812. doi:10.1002/recl.19290480808

CrossRef Full Text | Google Scholar

Jencks, W. P., and Carriuolo, J. (1960a). General Base Catalysis of the Aminolysis of Phenyl Acetate1. J. Am. Chem. Soc. 82, 675–681. doi:10.1021/ja01488a044

CrossRef Full Text | Google Scholar

Jencks, W. P., and Carriuolo, J. (1960b). Reactivity of Nucleophilic Reagents toward Esters. J. Am. Chem. Soc. 82, 1778–1786. doi:10.1021/ja01492a058

CrossRef Full Text | Google Scholar

Kamlet, M. J., Abboud, J. L. M., Abraham, M. H., and Taft, R. W. (1983). Linear Solvation Energy Relationships. 23. A Comprehensive Collection of the Solvatochromic Parameters, .pi.*, .alpha., and .beta., and Some Methods for Simplifying the Generalized Solvatochromic Equation. J. Org. Chem. 48, 2877–2887. doi:10.1021/jo00165a018

CrossRef Full Text | Google Scholar

Kirby, A. J., Dutta-Roy, N., da Silva, D., Goodman, J. M., Lima, M. F., Roussev, C. D., et al. (2005). Intramolecular General Acid Catalysis of Phosphate Transfer. Nucleophilic Attack by Oxyanions on the PO32- Group. J. Am. Chem. Soc. 127, 7033–7040. doi:10.1021/ja0502876

CrossRef Full Text | Google Scholar

Kirby, A. J., Manfredi, A. M., Souza, B. S. d., Medeiros, M., Priebe, J. P., Brandão, T. A. S., et al. (2008). Reactions of Alpha-Nucleophiles with a Model Phosphate Diester. Arkivoc 2009, 28–38. doi:10.3998/ark.5550190.0010.305

CrossRef Full Text | Google Scholar

Kirby, A. J., Tondo, D. W., Medeiros, M., Souza, B. S., Priebe, J. P., Lima, M. F., et al. (2009). Efficient Intramolecular General-Acid Catalysis of the Reactions of α-Effect Nucleophiles and Ammonia Oxide with a Phosphate Triester. J. Am. Chem. Soc. 131, 2023–2028. doi:10.1021/ja808746f

CrossRef Full Text | Google Scholar

Klopman, G., and Frierson, M. R. (1984). The Alpha-Effect. A Theoretical Study Incorporating Solvent Effects. Croat. Chem. Acta 57, 1411–1415.

Google Scholar

Kölmel, D. K., and Kool, E. T. (2017). Oximes and Hydrazones in Bioconjugation: Mechanism and Catalysis. Chem. Rev. 117, 10358–10376. doi:10.1021/acs.chemrev.7b00090

PubMed Abstract | CrossRef Full Text | Google Scholar

Kool, E. T., Crisalli, P., and Chan, K. M. (2014). Fast Alpha Nucleophiles: Structures that Undergo Rapid Hydrazone/Oxime Formation at Neutral pH. Org. Lett. 16, 1454–1457. doi:10.1021/ol500262y

PubMed Abstract | CrossRef Full Text | Google Scholar

Kwan, E. E., Zeng, Y., Besser, H. A., and Jacobsen, E. N. (2018). Concerted Nucleophilic Aromatic Substitutions. Nat. Chem 10, 917–923. doi:10.1038/s41557-018-0079-7

PubMed Abstract | CrossRef Full Text | Google Scholar

Lewis, G. N. (1923). Valence and the Structure of Atoms and Molecules. Am. Chem. Soc., Monograph Series. California: The Chemical Catalog Co., Inc.

Google Scholar

Liebman, J. F., Campbell, M. S., and Slayden, S. W. (1996). Thermochemistry of Amines, Nitroso Compounds, Nitro Compounds and Related Species. Chem. Amin. Nitroso, Nitro Relat. Groups, 337–378. doi:10.1002/047085720X.ch8

CrossRef Full Text | Google Scholar

Lowry, T. M. (1923). The Uniqueness of Hydrogen. J. Chem. Technol. Biotechnol. 42, 43–47. doi:10.1002/jctb.5000420302

CrossRef Full Text | Google Scholar

Makosza, M. (1993). Book Review: Nucleophilic Aromatic Displacement. The Influence of the Nitro Group.(Series: Organic Nitro Chemistry Series). By F. Terrier, Angew. Chem. Int. Ed. Engl. 32, 302–303. doi:10.1002/anie.199303022

CrossRef Full Text | Google Scholar

Neumann, C. N., Hooker, J. M., and Ritter, T. (2016). Concerted Nucleophilic Aromatic Substitution with 19F− and 18F−. Nature 534, 369–373. doi:10.1038/nature17667

PubMed Abstract | CrossRef Full Text | Google Scholar

Neumann, C. N., and Ritter, T. (2017). Facile C-F Bond Formation through a Concerted Nucleophilic Aromatic Substitution Mediated by the PhenoFluor Reagent. Acc. Chem. Res. 50, 2822–2833. doi:10.1021/acs.accounts.7b00413

PubMed Abstract | CrossRef Full Text | Google Scholar

Newington, I., Perez-Arlandis, J. M., and Welton, T. (2007). Ionic Liquids as Designer Solvents for Nucleophilic Aromatic Substitutions. Org. Lett. 9, 5247–5250. doi:10.1021/ol702435f

PubMed Abstract | CrossRef Full Text | Google Scholar

Nigst, T. A., Antipova, A., and Mayr, H. (2012). Nucleophilic Reactivities of Hydrazines and Amines: The Futile Search for the α-Effect in Hydrazine Reactivities. J. Org. Chem. 77, 8142–8155. doi:10.1021/jo301497g

CrossRef Full Text | Google Scholar

Nudelman, N. S. (2009). “SNAr Reactions of Amines in Aprotic Solvents,” in Patai’s Chemistry of Functional Groups, 2–5. doi:10.1002/9780470682531.pat0096

CrossRef Full Text | Google Scholar

Ormazábal-Toledo, R., Contreras, R., and Campodónico, P. R. (2013a). Reactivity Indices Profile: A Companion Tool of the Potential Energy Surface for the Analysis of Reaction Mechanisms. Nucleophilic Aromatic Substitution Reactions as Test Case. J. Org. Chem. 78, 1091–1097. doi:10.1021/jo3025048

CrossRef Full Text | Google Scholar

Ormazábal-Toledo, R., Contreras, R., Tapia, R. A., and Campodónico, P. R. (2013b). Specific Nucleophile-Electrophile Interactions in Nucleophilic Aromatic Substitutions. Org. Biomol. Chem. 11, 2302–2309. doi:10.1039/C3OB27450K

CrossRef Full Text | Google Scholar

Ormazabal-Toledo, R., Santos, J. G., Ríos, P., Castro, E. A., Campodónico, P. R., and Contreras, R. (2013). Hydrogen Bond Contribution to Preferential Solvation in SNAr Reactions. J. Phys. Chem. B 117, 5908–5915. doi:10.1021/jp4005295

CrossRef Full Text | Google Scholar

Ren, Y., and Yamataka, H. (2006). The α-Effect in Gas-phase SN2 Reactions Revisited. Org. Lett. 8, 119–121. doi:10.1021/ol0526930

PubMed Abstract | CrossRef Full Text | Google Scholar

Ren, Y., and Yamataka, H. (2007). The α-Effect in Gas-phase SN2 Reactions: Existence and the Origin of the Effect. J. Org. Chem. 72, 5660–5667. doi:10.1021/jo070650m

CrossRef Full Text | Google Scholar

Rohrbach, S., Murphy, J. A., and Tuttle, T. (2020). Computational Study on the Boundary between the Concerted and Stepwise Mechanism of Bimolecular SNAr Reactions. J. Am. Chem. Soc. 142, 14871–14876. doi:10.1021/jacs.0c01975

PubMed Abstract | CrossRef Full Text | Google Scholar

Sánchez, B., Calderón, C., Garrido, C., Contreras, R., and Campodónico, P. R. (2018a). Solvent Effect on a Model SNAr Reaction in Ionic Liquid/water Mixtures at Different Compositions. New J. Chem. 42, 9645–9650. doi:10.1039/C7NJ04820C

CrossRef Full Text | Google Scholar

Sánchez, B., Calderón, C., Tapia, R. A., Contreras, R., and Campodónico, P. R. (2018b). Activation of Electrophile/Nucleophile Pair by a Nucleophilic and Electrophilic Solvation in a SNAr Reaction. Front. Chem. 6, 509. doi:10.3389/fchem.2018.00509

CrossRef Full Text | Google Scholar

Stenlid, J. H., and Brinck, T. (2017). Nucleophilic Aromatic Substitution Reactions Described by the Local Electron Attachment Energy. J. Org. Chem. 82, 3072–3083. doi:10.1021/acs.joc.7b00059

PubMed Abstract | CrossRef Full Text | Google Scholar

Swager, T. M., and Wang, P. (2017). A Negotiation between Different Nucleophiles in SNAr Reactions. Synfacts 13, 0148. doi:10.1055/s-0036-1589929

CrossRef Full Text | Google Scholar

Terrier, F. (2013). The SNAr Reactions: Mechanistic Aspects. Mod. Nucleophilic Aromat. Substit., 231–361. doi:10.1002/9783527656141.ch1

CrossRef Full Text | Google Scholar

Um, I.-H., Chung, E.-K., and Lee, S.-M. (1998). An Unusual Ground-State Stabilization Effect and Origins of the Alpha-Effect in Aminolyses of Y-Substituted Phenyl X-Substituted Benzoates. Can. J. Chem. 76, 729–737. doi:10.1139/v98-043

CrossRef Full Text | Google Scholar

Um, I.-H., Hwang, S.-J., and Buncel, E. (2006). Solvent Effect on the α-Effect: Ground-State versus Transition-State Effects; a Combined Calorimetric and Kinetic Investigation. J. Org. Chem. 71, 915–920. doi:10.1021/jo051823f

CrossRef Full Text | Google Scholar

Um, I.-H., Im, L.-R., Kang, J.-S., Bursey, S. S., and Dust, J. M. (2012). Mechanistic Assessment of SNAr Displacement of Halides from 1-Halo-2,4-Dinitrobenzenes by Selected Primary and Secondary Amines: Brønsted and Mayr Analyses. J. Org. Chem. 77, 9738–9746. doi:10.1021/jo301862b

PubMed Abstract | CrossRef Full Text | Google Scholar

Um, I.-H., Min, S.-W., and Dust, J. M. (2007). Choice of Solvent (MeCN vs H2O) Decides Rate-Limiting Step in SNAr Aminolysis of 1-Fluoro-2,4-Dinitrobenzene with Secondary Amines: Importance of Brønsted-type Analysis in Acetonitrile. J. Org. Chem. 72, 8797–8803. doi:10.1021/jo701549h

PubMed Abstract | CrossRef Full Text | Google Scholar

Um, I.-H., Moon, H.-J., Shin, Y.-H., and Dust, J. M. (2018). Medium Effect on the α-effect for Nucleophilic Substitution Reactions of P-Nitrophenyl Acetate with Benzohydroxamates and M-Chlorophenoxide in DMSO-H2o Mixtures as Contrasts with MeCN-H2o Mixtures: Comparing Two Very Different Polar Aprotic Solvent Components. Can. J. Chem. 96, 922–928. doi:10.1139/cjc-2018-0103

CrossRef Full Text | Google Scholar

Von Bebenburg, W., and Thiele, K. (1970). Antipyretic 2-(phenylamino)- and 2-(pyridylamino)pyrimidines with an Amino or Amido Group in the 5-position. US3499898A. United States: United States Patent Office.

Google Scholar

Zingaretti, L., Boscatto, L., Chiacchiera, S. M., and Silber, J. J. (2003). Kinetics and Mechanism for the Reaction of 1-Chloro-2,4-Dinitrobenzene with N-Butylamine and Piperidine in AOT/n-hexane/water Reverse Micelles. Arkivoc 2003, 189–200. doi:10.3998/ark.5550190.0004.a19

CrossRef Full Text | Google Scholar

Keywords: SNAr reactions, mechanisms, hydrogen bond interaction, reactivity, brønsted type-plots

Citation: Campodónico PR, Tapia RA and Suárez-Rozas C (2022) How the Nature of an Alpha-Nucleophile Determines a Brønsted Type-Plot and Its Reaction Pathways. An Experimental Study. Front. Chem. 9:740161. doi: 10.3389/fchem.2021.740161

Received: 12 July 2021; Accepted: 30 December 2021;
Published: 02 February 2022.

Edited by:

Doo Soo Chung, Seoul National University, South Korea

Reviewed by:

Jose Ramón Mora, Universidad San Francisco de Quito, Ecuador
Jing Ma, Henan University, China

Copyright © 2022 Campodónico, Tapia and Suárez-Rozas. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Paola R. Campodónico, pcampodonico@udd.cl

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.