throbber
PCCP
`
`www.rsc.org/pccp
`
`Dynamic Article Links
`
`PAPER
`
`On the origin of the steric effectw
`Balazs Pinter,*a Tim Fievez,a F. Matthias Bickelhaupt,b Paul Geerlingsa and
`Frank De Proft*a
`
`Received 4th April 2012, Accepted 10th May 2012
`DOI: 10.1039/c2cp41090g
`
`A quantitative analysis of the steric effect of aliphatic groups was carried out from first principles.
`An intuitive framework is proposed that allows the separation and straightforward interpretation
`of two contributors to the steric effect: steric strain and steric shielding (hindrance). When a
`sterically demanding group is introduced near a reactive center, deformation of its reactive zone
`will occur. By quantifying this deformation, a convincing correlation was established with
`Taft’s steric parameters for groups of typical size, supporting the intuitive image of steric
`shielding; bulky groups slow down the reaction by limiting the accessibility of the reactive centre.
`On the other hand, the strong initial repulsion between the reactant and the substrate by means
`of the filled–filled orbital interaction results in the deformation of the substrate as well as a less
`stabilizing reaction zone, which are the manifestations of the steric strain. We thus conclude
`that both steric strain and steric hindrance can be derived from the Pauli repulsion evolving
`between the reactants in the course of the reaction.
`
`hydrolysis of substituted esters led Taft to define the total
` 
`steric effect, ES,10 of a substituent relative to a reference11 as
`ES ¼ log
`
`ð1Þ
`
`A
`
`k k
`
`0
`
`Introduction
`
`In spite of the early recognition of the importance of the steric
`effect1 and the spread of the concept throughout the chemical
`literature,2 by the beginning of the 1940s ‘‘steric hindrance. . .
`has become the last refuge of the puzzled organic chemist’’.3
`Reasons for very little progress in describing the steric effect in
`a quantitative manner during this period include scarce knowledge
`of reaction mechanisms, lack of proper experimental methods to
`determine the reaction rate and lack of reactions to separate
`electronic and steric effects and the general attempt to account
`for all chemical behaviours in terms of electronic effects due to
`the success of Stieglitz, Robinson, and Ingold’s electronic
`theory.4 The outstanding work of Ingold5 in the understanding
`of reaction mechanisms and that of Brown6 in developing
`techniques for the determination of kinetic and thermodynamic
`data eventually paved the way for the development of the first
`consistent quantitative scale of the steric effect introduced by
`Robert W. Taft.7
`Following the initial proposal of Ingold,8 the recognition
`that steric effects operate almost alone9 in acid catalysed
`
`a Eenheid Algemene Chemie (ALGC), Member of the QCMM
`VUB-UGent Alliance Research Group, Vrije Universiteit Brussel,
`Pleinlaan 2, Brussel, Belgium. E-mail: pbalazs@vub.ac.be,
`fdeprof@vub.ac.be; Fax: +32 2629 3317; Tel: +32 2629 3520
`b Department of Theoretical Chemistry and Amsterdam Center for
`Multiscale Modeling, Vrije Universiteit Amsterdam, De Boelelaan
`1083, Amsterdam, The Netherlands
`w Electronic supplementary information (ESI) available: Cartesian
`coordinates of the optimized structures,
`interaction potentials for
`1–22, transition state structure of 8-TS, correlation between XSE
`values and Taft constants. See DOI: 10.1039/c2cp41090g
`
`where k and k0 are the rate constants for the hydrolysis of the
`substituted and the reference esters, respectively. The average
`values of log(k/k0)A for five related reactions (vide infra) were
`used to reduce small specific effects and experimental errors in
`evaluating ES.12 Assuming that polar and steric effects dominate
`the reactivity of ester hydrolysis under basic conditions and
`that the steric effect is identical in the acid and base catalysed
`reaction, Taft suggested eqn (2) for evaluating the net polar
`effect of a substituent.7,13
`
`
`
` 
`
` 
`
`
`
`ð2Þ
`
`A
`
`k k
`
`0
`
` log
`
`B
`
`k k
`
`0
`
`s ¼ 1
`2:48
`
`log
`
`In this expression, the prefactor is merely introduced to put s*
`on the same scale as the well-known Hammett s values.14
`Shorter discussed the validity and limitations of Taft’s
`assumptions in detail.15 In spite of some criticisms of Taft’s
`steric parameters at various occasions16,17 and the development
`of modified,17 novel18 as well as corrected19 steric parameters,
`the ES scale nevertheless became and remained the standard
`quantitative measure of the steric effect in physical organic
`chemistry.
`Taft also played a pioneering role in the understanding of
`how the steric effect operates.20 His statements that ‘‘it is
`meaningless and deceptive to say that a molecule is ‘‘sterically
`hindered’’
`from a general reactivity point of view’’ and
`
`9846
`
`Phys. Chem. Chem. Phys., 2012, 14, 9846–9854
`
`This journal is c the Owner Societies 2012
`
`Cite this: Phys.Chem.Chem.Phys., 2012, 14, 9846–9854
`
`Published on 10 May 2012. Downloaded by Reprints Desk on 2/19/2020 9:38:57 PM.
`
`View Article Online
`
` / Journal Homepage
`
` / Table of Contents for this issue
`
`SYNGENTA EXHIBIT 1017
`Syngenta v. FMC, PGR2020-00028
`
`

`

`‘‘changes in steric interactions between reactant and transition
`states are the only factors that affect rates’’ clearly highlight
`his interpretation of the steric effect within the framework of
`transition state theory.20 Accordingly, the steric contributions
`to the activation free energy were analysed and classified into
`potential- and kinetic-energy factors. It is straightforward to
`see how the former, also called steric strain, operates: since the
`coordination number of the carbonyl carbon changes from
`three to four when going from the reactant to the transition
`state during the acid-catalysed ester hydrolysis, the presence of
`substituents of increasing size near the reactive centre results in
`an increasing potential energy or steric strain due to the
`repulsion between non-bonded atoms. The kinetic-energy
`steric effect was associated with the change in steric hindrance
`of internal motions during the activation process.
`This effect is analogous to the entropy contribution to the
`activation free energy described in the principle of Prince and
`Hammett;21 when going from the reactant to the transition
`state, bulkier functional groups will give rise to more entropy
`loss as compared to smaller groups. In-depth investigation of the
`measured kinetic parameters revealed that this steric hindrance of
`internal motions dominates the total steric energy of activation
`for substituents of small (e.g. –CH3, i-butyl) and medium sizes
`(e.g. –CH2(cyclohexyl), –CH(C6H5)2). Introduction of even larger
`groups (i.e. beyond the g position relative to the carbonyl carbon
`atom, such as e.g. –CH(tert-butyl)(CH3), –C(C2H5)3) leads to an
`increasing importance of steric strain, but never without an
`accompanying increased steric hindrance of motions.22 This more
`negative activation entropy due to the introduction of bulkier
`groups DS#
`steric can be also linked to the steric factor P of collision
`theory through the expression P ¼ eDS#
`steric . This provides a
`natural bridge between transition state theory and collision
`theory concerning the steric effect.23
`Numerous attempts have been made to relate the ES constants
`to topological indices and geometrical, structural or physical
`parameters of substituents. From the excellent correlation found
`between ES values and van der Waals radii of groups (r$ as
`depicted in Scheme 1a in the case of a –CX3 group) Charton
`concluded24 that ES values are indeed free of inductive and
`resonance effects and, consequently, are pure measures of steric
`effects. However, this work used different sets, such as –CH2X,
`–CHX2 and –CX3, each with only five to seven substituents and
`with small overall ES ranges, which might be insufficient to
`clearly separate steric and electronic effects.25 Another wide-
`spread principle, Newman’s rule of six,26 which states that the
`number of atoms in the sixth position from the carbonyl oxygen
`(see Scheme 1b) determines the steric effect, was established by
`
`Scheme 1 Charton’s size-related metrics for –CX3 type groups (side
`view and top view) (a), Newman’s rule of six (b) and steric congestion
`as the inverse of accessibility of the carbonyl carbon for the approaching
`reactant R (c).
`
`qualitatively testing several reactions.27 This approach how-
`ever is evidently limited to b substitution of alkyl groups.
`These empirical methods have the important advantage of
`simplicity and are certainly useful for rough estimation of the
`steric effect, however, they do not provide insight into the
`nature of the steric hindrance, which is crucial in developing a
`universal concept of this quantity.
`The conceptually more elaborate but still purely geometrical
`method of Wipke and Gund28 was intended to predict the
`steric effect in the reactant state of the substrate independently
`of the structure of the transition state and reaction partner.
`Their steric congestion (Cxa, Scheme 1c), evaluated by eqn (3)
`and (4), describes the accessibility of the reactive carbonyl
`carbon (C1) for an approaching reactant R. It is clearly an
`approximation for the cross-section in collision theory. As
`such, it covers at least a part of the entropy factor although it
`is not straightforward how to interpret Cxa in Taft’s potential-
`and kinetic-energy steric effect framework (vide supra). The
`calculated congestions gave satisfactory correlation with the
`stereoselectivity of nucleophilic addition to 52 ketones; how-
`ever, they did not correlate as well with the absolute rate of the
`reaction.
`
`Axa(i) = 2pr2(1 cosy)
`
`Cxa = SiCxa(i) = Si(1/Axa(i))
`
`(3)
`
`(4)
`
`In molecular mechanics based methods, steric effects are
`treated as the sum of bond length and angle deformation
`(Baeyer strain), torsional eclipsing energy (Pitzer strain) and
`non-bonded interactions.29 The key idea of these methods is
`that the deviation in steric strain (DSE) when going from the
`reactant to the transition state equals the activation enthalpy
`(DH) and was established for various solvolysis reactions by
`obtaining linear correlation with the corresponding experi-
`mental rates.30 No such relationship, however, could be
`established for SN2 type reactions.31 The linear expression
`log(k/k0) = 0.340 0.789SE found for acid-catalyzed ester
`hydrolysis supports, then again, that ES values do in fact
`provide an exclusive measure of the steric effect.32
`Despite the developments in computational chemistry codes
`and supercomputers, the times where ‘‘. . .qualitative application of
`the theories of steric effects, aided by rapid computer calculations
`of molecular and transition state structures. . .’’ are still yet to
`come.33 Instead, practicing mechanistic computational chemists
`often regard steric effects as a universal redeeming rationale for
`phenomena that cannot be interpreted on electrostatic and orbital
`interaction origins. This is actually not inappropriate concerning
`the more or less obscure concept of the steric effect, which in
`practice embodies the views of strain, hindrance of internal
`motions, kinetic trapping, repulsion between non-bonded atoms,
`etc. In other words, as expressed by Ferna´ ndez, Frenking and
`Uggerud, the operationally defined steric effect lacks ‘‘rigorous
`mathematical accuracy’’.34 Unfortunately, the situation is more
`complicated since the minimal energy difference in activation
`barriers, sometimes as small as 2–3 kcal mol1 (that makes the
`difference between inhibited (slow) and allowed (fast) reaction)
`that we wish to capture and interpret, is far beyond the accuracy of
`most of the applied methods. No new concepts have completely
`resolved the post-Babelian conditions concerning the steric effect;
`
`This journal is c the Owner Societies 2012
`
`Phys. Chem. Chem. Phys., 2012, 14, 9846–9854
`
`9847
`
`Published on 10 May 2012. Downloaded by Reprints Desk on 2/19/2020 9:38:57 PM.
`
`View Article Online
`
`

`

`recently Weizsa¨ cker kinetic energy,35 Fisher information,36
`steric energy37 as well as various isodesmic38 reactions were
`proposed as a measure of the steric effect. Moreover, a very
`recent electron density based approach introduced by Yang
`and co-workers allows the visualization of, besides other
`fundamental properties, steric clashes in small molecules,
`molecular complexes and solids.39
`Being the only source of net repulsive interactions between
`molecular fragments, Pauli repulsion has been associated with
`the steric interaction for some time now.40 Very recently its
`quantification within the framework of the activation-strain
`model has shed light on the actions of the steric effect in
`bimolecular SN2 reactions41 and oxidative addition/insertion
`processes.42 Going from hydrogen to methyl substituents the
`larger steric effect was manifested in higher ‘‘activation’’
`strain, the energy associated with deforming the reactants
`from their equilibrium geometry, as well as in stronger Pauli
`repulsion between fragments. It turns out that the excess
`‘‘activation’’ strain originates from and its magnitude depends
`on the Pauli repulsion too: a strong initial Pauli repulsion
`results in an additional reactant deformation that actually
`relieves part of the inducing steric interaction.43 This concept
`provides a rational interpretation of the steric strain, the steric
`effect that is manifested in the potential energy, but however,
`does not account for and cannot describe the contribution of
`steric shielding.
`In this contribution we interpret and quantify the two
`actions of the steric effect: steric shielding and steric strain.
`We demonstrate that for groups of typical size, the steric
`effect, manifested in rate retardation of ester hydrolysis and
`represented by Taft constants, is dominated by shielding rather
`than strain. As a proof of concept we provide quantitative
`justifications for the intuitive image that bulky groups introduced
`near the reactive centre slow down the reaction by limiting the
`accessibility of the reactive centre. Our framework allows for
`straightforward and intuitive distinction of these conceptually
`different actions of the steric effect.
`
`Computational details
`
`In order to scrutinize the steric effect, we focus on one of its
`defining reactions, the acid catalysed hydrolysis of esters. To
`this end, protonated esters 1–19 and 20–22 were optimized at
`B3LYP44/aug-cc-pVTZ45 and B3LYP/aug-cc-pVDZ levels,
`respectively, using Gaussian0346 software.
`Using equilibrium geometries, interaction energy calculations
`for the approach of the model nucleophile were carried out with
`the PBE47 functional in combination with a TZ2P basis set using
`the Amsterdam Density Functional (ADF) program package.48
`To evaluate the two contributions of the steric effect for groups
`1–22, interaction potentials were determined on a 4  4 A˚ square
`grid centred above C1, at 2 A˚
`from the substrate, using 441
`points (ca. 28 points A˚ 2). The distance of 2 A˚ between the
`fragments is chosen to allow the description of the steric effect
`in the initial stages of the bond formation process. Note that
`specifying the electron configuration of the fragments ensures
`the required ionic interaction of the reactants, i.e. the alter-
`native association of the reactants as a radical pair, which
`might be appropriate at larger distances, is thus eliminated
`
`from the description. For details of our protocol of calculating
`interaction energy potentials see ref. 49.
`The narrowing of the reactive area was quantified using
`an exclusion method. We defined that the reference system
`(R = CH3) has a circular reactive area with a radius of 1 A˚ at
`2 A˚
`from the reactive centre. By investigating the same circle
`centred above the reactive carbon for our test series we can
`determine what percentage of this area is available for attack
`for a given substituent. By setting a threshold limit for the
`Pauli repulsion we can easily decide whether a point in the
`circle is a part of the reactive zone or the incoming nucleophile
`would feel too strong repulsion to approach this point. Based
`on the analysis below, we applied a standard limit value of
`100 kcal mol1; if the Pauli repulsion is lower than this value in
`a given point then that point is available for attack whereas for
`higher values it is excluded from the reactive zone.
`Interaction energy contour plots were generated with Mathe-
`1. The distance
`matica and all contour lines are given in kcal mol
`1 for
`between contour lines is consistently fixed to 10 kcal mol
`
`DEPauli, DVelst and DEoi and to 5 kcal mol1 for DEnoi and DEint.
`The color-coding is chosen such that dark regions correspond to
`the most stable zones and the light regions to the least stable areas
`in the plot (e.g. in Fig. 2). Areas where the reactant is too close
`to the substrate are indicated in white.
`
`Results
`
`To quantify the steric effect of aliphatic groups, we selected and
`investigated a set of protonated esters 1–22 (Scheme 2) that are
`the active species of the rate determining step in acid catalyzed
`ester hydrolysis, one of the defining reactions of Taft’s constant.
`For such a reaction, as it was demonstrated (vide supra) groups
`introduced close to the reactive center, i.e. at the carbonyl carbon
`atom, contribute to the increase of the activation barrier through
`the enthalpy as well as the entropy. These contributions change
`proportionally only in a few specific cases50 and hence, any
`model intended to describe the steric effect should properly
`consider and characterize the two contributions separately. It
`implies, as was also noted by others, that single parameter
`models are de facto bogus. Consequently, our model will use
`two parameters; the area of the reactive zone and the average
`Pauli repulsion within this area.
`
`Scheme 2 Investigated protonated methylesters 1–22 and numbering
`of the relevant atoms.
`
`9848
`
`Phys. Chem. Chem. Phys., 2012, 14, 9846–9854
`
`This journal is c the Owner Societies 2012
`
`Published on 10 May 2012. Downloaded by Reprints Desk on 2/19/2020 9:38:57 PM.
`
`View Article Online
`
`

`

`The area of the reactive zone, the surface through which the
`reactant can approach the reactive site of the substrate, is
`introduced to measure the accessibility of the reactive centre. It
`resembles the reactive cross-section s* in collision theory, defined
`as s* = Ps, where P is the steric factor ðP ¼ eDS#
`stericÞ and s is
`the collision cross-section.23 Accordingly, we define the steric
`shielding as the process by which sterically demanding groups
`limit the accessibility of the reactive centre.
`Bulky groups do not only limit the accessibility of the
`reactive centre but also exert considerable repulsion against
`the incoming reactant due to the filled–filled orbital interaction.
`Such repulsion induces the deformation of the substrate on the
`one hand and results in a less stabilizing interaction energy on
`the other. Taking into account that resonance and inductive
`effects are minor and do not vary with R in the investigated
`reaction, one can expect that orbital and electrostatic interac-
`tions between the incoming reactant and the reactive centre are
`nearly constant for varying R. Thus, the total interaction energy
`along the reaction coordinate becomes less stabilizing with
`increasing group size solely because of the stronger Pauli
`repulsion between the substrate and reactant. Since both the
`deformation and the change in the interaction energy are
`manifested as an increase of potential energy, their effect is
`considered to be the steric strain and quantified by the average
`Pauli repulsion in the reactive zone.
`Our methodology for evaluating the steric shielding as well
`as the steric strain of alkyl groups is straightforward; as
`illustrated in Fig. 1, a model nucleophile (F) is moved along
`a rectangular grid parallel to the skeleton, defined by the O1,
`C1 and O2 atoms, of the protonated ester [R–C(OH)–OCH3]+
`and the interaction energy between the substrate and the nucleo-
`phile is determined in every point of the grid. The origin of the
`grid, [0, 0], was set to C1 whereas axes x and y were chosen to be
`parallel with and perpendicular to the C1–O2 bond, respectively.
`The fluoride ion was chosen as the model nucleophile to avoid
`complications of directional dependence. Its electrostatic inter-
`action with the protonated substrate is considered to produce a
`linearly scaled background compared to a neutral nucleophile. In
`every scanned point along the grid, the interaction energy DEint
`was decomposed into Pauli, electrostatic and orbital interaction
`terms using the Ziegler–Rauk decomposition scheme.51 Such a
`protocol enables us to visualize in-plane contour plots for
`these fundamental properties (DEPauli, DVelst and DEoi) and
`for
`two derived quantities,
`the non-orbital
`interaction
`
`Fig. 1 Schematic representation of our approach to quantify the
`steric effect of aliphatic groups; scans of the interaction energy
`between the protonated ester and the model nucleophile are performed
`along a rectangular grid at 2 A˚
`from the substrate. The green area is
`used to determine the steric effect of the substituent. Only plane1 is
`shown in the case of R = tert-butyl for simplicity.
`
`(DEnoi = DEPauli + DVelst) and total
`interaction energies
`(DEint = DEPauli + DVelst + DEoi), as functions of the relative
`position of the nucleophile to the ester. (Note: ‘‘non-orbital
`interaction’’ refers to all interactions except for the stabilizing
`orbital interactions.) As we will demonstrate, the interaction
`energy between the nucleophile and the substrate, which we
`will term the ‘‘interaction energy potential’’, contains enough
`information to quantify the steric effect of alkyl substituents in
`the molecule.
`interaction energy potential (DEint) calculated
`The total
`for the methyl substituted protonated ester (Fig. 2) clearly
`identifies the reactive zone, an attractive area around the
`minimal potential at position [0, 0],
`through which the
`nucleophile attacks the target carbon centre. The other attrac-
`tive zone around [(1/2), 2] orients the incoming nucleophile
`to an unreactive site of the substrate probably resulting in a
`non-reactive collision.
`DEpauli shows that the –OCH3 moiety and the group R exert
`considerable Pauli repulsion when the reactant approaches
`from their directions. The effect of the latter substituent is
`‘delocalized’ in the direction of the central carbon meaning
`that the exerted repulsion considerably affects the potential
`above and below the reactive carbon, which is the footprint of
`the steric effect.
`The electrostatic potential (DVelst) closely resembles the
`Pauli repulsion potential, but with opposite sign; due to slight
`electron deficient character of the out-of-plane hydrogens, the
`incoming nucleophile experiences the strongest electrostatic
`attraction in the proximity of the methyl groups. Also, the
`effect of partially localized positive charge is manifested in the
`rather attractive electrostatics above the central carbon atom.
`As shown in the DEoi potential, orbital interaction plays an
`important role in controlling the approach of the incoming
`nucleophile:
`the orbital
`interaction attraction above the
`reactive carbon is dominant enough to modify the appearance
`of DEnoi resulting in a preferred almost perpendicular approach
`to the carbon, in agreement with the calculated transition state
`geometry (Fig. S1, ESIw).52 The large distance of 1.74 A˚
`calculated between the reactive carbon and the oxygen atom
`of the incoming water molecule and the slight pyramidalization
`around C1 indicate an early TS for the hydrolysis. Such small
`geometrical distortion of the substrate when going from the
`reactant to the TS ensures that the steric effect can be described
`during the initial stages of the reaction.
`Scanning the interaction energy in planes at different distances
`from the molecule allows us to construct the three-dimensional
`picture of the ‘reactive channel’, the energetically preferred region
`above the reactive carbon that provides the optimal approach for
`the nucleophile. Such a channel with a typical conical shape is
`illustrated in Fig. 3 with interaction potential isosurface values of
`(a) 40 kcal mol
`1 and (b) 130 kcal mol
`1; the former also
`shows a valley above the central carbon allowing the nucleophile
`to get closer and interact with the carbon atom. Note that the
`reactive channel shown in Fig. 3b closely resembles the
`hypothetical cone in Wipke’s steric congestion method.28
`Changing the substituent on the central carbon (C1,
`Scheme 1) atom has a crucial impact on the shape of reactive
`channels. For instance, bulky groups that bend out of the
`plane in the proximity of C1 exert strong Pauli repulsion in the
`
`This journal is c the Owner Societies 2012
`
`Phys. Chem. Chem. Phys., 2012, 14, 9846–9854
`
`9849
`
`Published on 10 May 2012. Downloaded by Reprints Desk on 2/19/2020 9:38:57 PM.
`
`View Article Online
`
`

`

`Fig. 3 Shape of the reactive channel of the methyl substituted protonated
`ester with (a) 40 kcal mol1 and (b) 130 kcal mol1 isosurface values.
`
`Fig. 4 Interaction energy (top) and Pauli repulsion (bottom) potentials
`above the reactive carbon center for methyl (1), axial cyclohexyl
`(110, plane2) and tert-butyl (8, plane2) substituents.
`
`how the area and shape of the reactive zone, the cross-section
`of the reactive channel, change as a function of the substituent.
`The reference system (1: R = CH3) almost has a circular
`reactive zone with radius of about 1 A˚ as can be seen from the
`DEint plot. This circular shape deforms to oval and to a
`semicircle when going from methyl to cyclohexyl (110) and
`to tert-butyl substituents (8).
`Comparison of the corresponding total interaction energy
`and the Pauli repulsion potentials (Fig. 4) clearly shows
`that the deformation of the reactive zone originates from
`the overwhelming Pauli repulsion in the proximity of the
`substituent. As a rule of thumb, we found that when the Pauli
`term reaches a value of about 100 kcal mol1,
`it also
`dominates the overall appearance of the total
`interaction
`energy map. We now conjecture that the two important
`parameters for describing the steric effect provided by these
`calculations are the area of the reactive zone and the inter-
`action energy in this zone. To evaluate the steric effect of the
`functional groups, we propose that the size of the reactive
`area describes the steric shielding, whereas the average Pauli
`repulsion in the reactive area describes the strain energy
`contribution to the steric effect.
`For a set of alkyl substituted protonated esters, Table 1
`summarizes the size of the reactive zone in points (N) through
`which the reactant can approach the reactive carbon. In the
`absence of a mirror plane the two sides of the molecule from
`which the nucleophile can approach are not identical and need
`
`Fig. 2 Contour plot of the Pauli repulsion (DEPauli), electrostatic inter-
`action (DVelst), orbital interaction (DEoi), non-orbital interaction (DEnoi)
`and total interaction energy (DEint) potentials along a 6 A˚  5 A˚ grid for 1
`(R = CH3) in a plane that is 2 A˚ above the molecular plane.
`
`neighbourhood of the reactive carbon resulting in a less
`attractive and deformed reactive channel. Fig. 4 illustrates
`
`9850
`
`Phys. Chem. Chem. Phys., 2012, 14, 9846–9854
`
`This journal is c the Owner Societies 2012
`
`Published on 10 May 2012. Downloaded by Reprints Desk on 2/19/2020 9:38:57 PM.
`
`View Article Online
`
`

`

`Table 1 Taft constant (ES),16 number of reactive zone points in plane1 (N1) and plane2 (N2), average number of reactive zone points (Navg),
`average Pauli repulsion in the reactive zone of plane1 (hDEPaulii(1)) and plane2 (hDEPaulii(2)) in kcal mol1 and logarithm of relative reactive zone
`
`
`size log(Navg(R)/Navg(1)) for the investigated protonated esters
`log NavgðRÞ
`Navgð1Þ
`
`hDEPaulii(1)
`
`hDEPaulii(2)
`
`R
`
`ES
`
`N1
`
`N2
`
`Navg
`
`1
`2
`3
`4
`5
`6
`7
`8
`9
`10
`11
`110
`12
`13
`14
`140
`15
`16
`17
`18
`19
`20
`21
`210
`22
`
`–CH3
`–CH2CH3
`n-Propyl
`n-Butyl
`n-Pentyl
`i-Propyl
`i-Butyl
`tert-Butyl
`Cyclobutyl
`Cyclopentyl
`Cyclohexyleq
`Cyclohexylax
`Cycloheptyl
`–CH2C(CH3)3
`–(CH2)2CH(CH3)2
`–(CH2)2CH(CH3)2(2)
`–CH2Ph
`–C(CH2CH3)3
`–CH(CH3)(CH2CH3)
`–CH(CH3)(Ph)
`–CH(CH2CH3)2
`–(1-CH3–c-hexyl)
`–C(CH3)2(c-hexyl)
`–C(CH3)2(c-hexyl)(2)
`–CHPh2
`
`0.00
`0.07
`0.36
`0.39
`0.40
`0.47
`0.93
`1.54
`0.06
`0.51
`0.79
`0.79
`1.10
`1.74
`0.35
`0.35
`0.37
`3.80
`1.13
`1.19
`1.98
`2.03
`2.49
`2.49
`1.76
`
`79
`77
`60
`72
`72
`72
`9
`29
`81
`78
`43
`67
`67
`6
`77
`4
`73
`30
`24
`27
`19
`11
`48
`26
`2
`
`79
`77
`72
`63
`63
`25
`78
`53
`37
`20
`41
`20
`5
`81
`79
`67
`75
`15
`64
`70
`31
`47
`0
`22
`67
`
`79
`77
`66
`67.5
`67.5
`48.5
`43.5
`41
`59
`49
`42
`43.5
`36
`43.5
`78
`35.5
`74
`22.5
`44
`48.5
`25
`29
`24
`24
`34.5
`
`69.20
`69.64
`72.97
`71.43
`71.05
`71.40
`80.51
`74.41
`65.87
`68.19
`75.33
`71.36
`72.02
`83.19
`70.56
`88.18
`71.98
`73.73
`75.47
`74.35
`77.10
`84.44
`—
`77.32
`93.74
`
`69.20
`69.63
`71.82
`73.61
`73.63
`75.42
`69.13
`75.47
`76.17
`76.28
`76.46
`77.34
`87.16
`69.23
`70.18
`73.72
`71.08
`82.83
`71.41
`72.90
`78.61
`77.77
`76.50
`81.39
`73.03
`
`0.00
`0.01
`0.08
`0.07
`0.07
`0.21
`0.26
`0.28
`0.13
`0.21
`0.27
`0.26
`0.34
`0.26
`0.01
`0.35
`0.03
`0.55
`0.25
`0.21
`0.50
`0.44
`0.52
`0.52
`0.36
`
`to be evaluated separately (plane1 (N1) and plane2 (N2)).
`At the applied resolution of 28 points A˚ 2 the default analysed
`circle consists of 81 points. As expected, the highest average
`number of points occurs for the methyl substituted reference
`system (Navg(1) = 79). The advantageous orientation of
`cyclobutyl and –CH2C(CH3)3 substituents results in one
`completely unhindered side in 9 and 13, respectively, although,
`the overall accessibility of C1 is worse in these systems,
`Navg(9) = 59 and Navg(13) = 43.5, than for 1. In good
`agreement with the experimental measurement the predicted
`most shielded carbon centre appears in system 16 having the
`bulky –C(CH2CH3)3 substituent. Introduction of two phenyl
`rings on C2, as expected, has a prominent influence on the
`shielding; plane1 of the –CHPh2 substituted protonated ester
`(22) is more shielded (N1(22) = 2) than either side of the
`–C(CH2CH3)3 substituted derivative (16).
`The average Pauli repulsion calculated in the reactive zone
`(hDEPaulii(1) and hDEPaulii(2)) varies between 66 kcal mol
`1
`1 (R = CHPh2
`(R = cyclobutyl (9), plane1) and 94 kcal mol
`(22), plane1). No unambiguous direct correlation can be
`recognized between the Taft constants and this average
`hDEPaulii, however, the statement that ‘‘the bigger the reactive
`zone the smaller the repulsion experienced’’ is valid in most of
`the cases. For example, in the plane of 12 and 13, several
`accessible points (B5) are associated with relatively high
`average repulsion of about 85 kcal mol1 whereas the reactive
`zone of 67 points in 110 (plane1) and 12 (plane1) has a
`hDEPaulii(1) of only B72 kcal mol1. The wide range of almost
`1 for the average hDEPaulii, which would result in
`30 kcal mol
`an unrealistic 4.6  1021-fold retardation of the reaction rate
`at 30 1C, shows that the predicted strain effect has to be scaled
`down. For this sole purpose we introduce the parameter a,
`with an expected value of 0.05–0.1, to put the effect of
`
`hDEPaulii on the same scale as for the shielding effect described
`by the change in the reactive zone area.
`We finally propose to estimate the steric effect XSE of
`group R as
`
` 
`XSEðRÞ ¼ log
`
`kX
`R
`kX
`1
`
`ð5aÞ
`
`RT
`
`RT
`
`þ N2ðRÞeahDEPauliðRÞið2Þ
`R ¼ N1ðRÞeahDEPauliðRÞið1Þ
` kR ð5bÞ
`kX
`where N1(1) = N2(1) = 79 and hDEPauli(1)i(1) = hDEPauli(1)i(2) =
`69.2 kcal mol1 are the values corresponding to the methyl
`substituent. To test the reliability of our method and validity of
`our assumptions, XSE values for various a parameters were plotted
`against the Taft constants and their correlation was determined.
`Using an a value of 0.01, a satisfactory agreement (R2 = 0.82) was
`found between calculated XSE and experimental ES parameters
`(Fig. S2, ESIw).
`Importantly, the small a value indicates that the exponential
`term in eqn (5a) and (5b), i.e. the steric strain in our framework,
`affects the correlation insignificantly. Accordingly, we investigated
`the foreshadowed relationship between the relative size of the
`reactive zone (log(Navg(R)/Navg(1))) and Taft constants; as can be
`seen in Fig. 5, an acceptable correlation with R2 of 0.82 indeed
`exists between the relative reactive zone size for protonated esters
`1–22 and steric constants of the corresponding aliphatic groups.
`One can interpret the latter correlation and the predicted small a
`parameter as an indication of negligible steric strain contribution
`to the decrease of the reaction rate for the majority of the
`investigated substrates as an a value of 0.01 results in a predicted
`maximal strain contribution of 0.3 kcal mol1. Such interpretation
`is indeed partially valid; Taft had shown that many of the
`investigated substrates exhibit a considerable entropy contribution
`
`This journal is c the Owner Societies 2012
`
`Phys. Chem. Chem. Phys., 2012, 14, 9846–9854
`
`9851
`
`Published on 10 May 2012. Downloaded by Reprints Desk on 2/19/2020 9:38:57 PM.
`
`View Article Online
`
`

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket