57
doi.org/10.26434/chemrxiv.12758498.v1 Machine Learning Meets Mechanistic Modelling for Accurate Prediction of Experimental Activation Energies Kjell Jorner, Tore Brinck, Per-Ola Norrby, David Buttar Submitted date: 04/08/2020 Posted date: 05/08/2020 Licence: CC BY-NC-ND 4.0 Citation information: Jorner, Kjell; Brinck, Tore; Norrby, Per-Ola; Buttar, David (2020): Machine Learning Meets Mechanistic Modelling for Accurate Prediction of Experimental Activation Energies. ChemRxiv. Preprint. https://doi.org/10.26434/chemrxiv.12758498.v1 Accurate prediction of chemical reactions in solution is challenging for current state-of-the-art approaches based on transition state modelling with density functional theory. Models based on machine learning have emerged as a promising alternative to address these problems, but these models currently lack the precision to give crucial information on the magnitude of barrier heights, influence of solvents and catalysts and extent of regio- and chemoselectivity. Here, we construct hybrid models which combine the traditional transition state modelling and machine learning to accurately predict reaction barriers. We train a Gaussian Process Regression model to reproduce high-quality experimental kinetic data for the nucleophilic aromatic substitution reaction and use it to predict barriers with a mean absolute error of 0.77 kcal/mol for an external test set. The model was further validated on regio- and chemoselectivity prediction on patent reaction data and achieved a competitive top-1 accuracy of 86%, despite not being trained explicitly for this task. Importantly, the model gives error bars for its predictions that can be used for risk assessment by the end user. Hybrid models emerge as the preferred alternative for accurate reaction prediction in the very common low-data situation where only 100–150 rate constants are available for a reaction class. With recent advances in deep learning for quickly predicting barriers and transition state geometries from density functional theory, we envision that hybrid models will soon become a standard alternative to complement current machine learning approaches based on ground-state physical organic descriptors or structural information such as molecular graphs or fingerprints. File list (3) download file view on ChemRxiv Manuscript.pdf (1.45 MiB) download file view on ChemRxiv ESI.pdf (1.48 MiB) download file view on ChemRxiv ESI data.zip (95.01 MiB)

Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

doi.org/10.26434/chemrxiv.12758498.v1

Machine Learning Meets Mechanistic Modelling for Accurate Predictionof Experimental Activation EnergiesKjell Jorner, Tore Brinck, Per-Ola Norrby, David Buttar

Submitted date: 04/08/2020 • Posted date: 05/08/2020Licence: CC BY-NC-ND 4.0Citation information: Jorner, Kjell; Brinck, Tore; Norrby, Per-Ola; Buttar, David (2020): Machine LearningMeets Mechanistic Modelling for Accurate Prediction of Experimental Activation Energies. ChemRxiv.Preprint. https://doi.org/10.26434/chemrxiv.12758498.v1

Accurate prediction of chemical reactions in solution is challenging for current state-of-the-art approachesbased on transition state modelling with density functional theory. Models based on machine learning haveemerged as a promising alternative to address these problems, but these models currently lack the precisionto give crucial information on the magnitude of barrier heights, influence of solvents and catalysts and extentof regio- and chemoselectivity. Here, we construct hybrid models which combine the traditional transition statemodelling and machine learning to accurately predict reaction barriers. We train a Gaussian ProcessRegression model to reproduce high-quality experimental kinetic data for the nucleophilic aromaticsubstitution reaction and use it to predict barriers with a mean absolute error of 0.77 kcal/mol for an externaltest set. The model was further validated on regio- and chemoselectivity prediction on patent reaction dataand achieved a competitive top-1 accuracy of 86%, despite not being trained explicitly for this task.Importantly, the model gives error bars for its predictions that can be used for risk assessment by the enduser. Hybrid models emerge as the preferred alternative for accurate reaction prediction in the very commonlow-data situation where only 100–150 rate constants are available for a reaction class. With recent advancesin deep learning for quickly predicting barriers and transition state geometries from density functional theory,we envision that hybrid models will soon become a standard alternative to complement current machinelearning approaches based on ground-state physical organic descriptors or structural information such asmolecular graphs or fingerprints.

File list (3)

download fileview on ChemRxivManuscript.pdf (1.45 MiB)

download fileview on ChemRxivESI.pdf (1.48 MiB)

download fileview on ChemRxivESI data.zip (95.01 MiB)

Page 2: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

1

Machine learning meets mechanistic modelling for accurate prediction of experimental activation energies Kjell Jorner,1 Tore Brinck,2 Per-Ola Norrby,3 David Buttar1 1 Early Chemical Development, Pharmaceutical Sciences, R&D, AstraZeneca, Macclesfield, United Kingdom 2 Applied Physical Chemistry, Department of Chemistry, CBH, KTH Royal Institute of Technology, Stockholm, Sweden 3 Data Science & Modelling, Pharmaceutical Sciences, R&D, AstraZeneca, Gothenburg, Sweden

Abstract Accurate prediction of chemical reactions in solution is challenging for current state-of-the-art

approaches based on transition state modelling with density functional theory. Models based on

machine learning have emerged as a promising alternative to address these problems, but these models

currently lack the precision to give crucial information on the magnitude of barrier heights, influence of

solvents and catalysts and extent of regio- and chemoselectivity. Here, we construct hybrid models

which combine the traditional transition state modelling and machine learning to accurately predict

reaction barriers. We train a Gaussian Process Regression model to reproduce high-quality experimental

kinetic data for the nucleophilic aromatic substitution reaction and use it to predict barriers with a mean

absolute error of 0.77 kcal/mol for an external test set. The model was further validated on regio- and

chemoselectivity prediction on patent reaction data and achieved a competitive top-1 accuracy of 86%,

despite not being trained explicitly for this task. Importantly, the model gives error bars for its

predictions that can be used for risk assessment by the end user. Hybrid models emerge as the preferred

alternative for accurate reaction prediction in the very common low-data situation where only 100–150

rate constants are available for a reaction class. With recent advances in deep learning for quickly

predicting barriers and transition state geometries from density functional theory, we envision that

hybrid models will soon become a standard alternative to complement current machine learning

approaches based on ground-state physical organic descriptors or structural information such as

molecular graphs or fingerprints.

Introduction Accurate prediction of chemical reactions is an important goal both in academic and industrial

research.1–3 Recently, machine learning approaches have had tremendous success in quantitative

prediction of reaction yields based on data from high-throughput experimentation4,5 and

enantioselectivities based on carefully selected universal training sets.6 At the same time, traditional

quantitative structure-reactivity relationship (QSRR) methods based on linear regression have seen a

renaissance with interpretable, holistic models that can generalize across reaction types.7 In parallel with

these developments of quantitative prediction methods, deep learning models trained on reaction

databases containing millions of patent and literature data have made quick qualitative yes/no feasibility

prediction routine for almost any reaction type.8

In the pharmaceutical industry, prediction tools have great potential to accelerate synthesis of

prospective drugs (Figure 1a).9 Quick prediction is essential in the discovery phase, especially within the

context of automation and rapid synthesis of a multitude of candidates for initial activity screening.3,10,11

In these circumstances, a simple yes/no as provided by classification models is usually sufficient. More

accurate prediction is necessary in the later drug development process, where the synthesis route and

formulation of one or a few promising drug candidates is optimized. Here, regression models that give

the reaction activation energy can be used to predict both absolute reactivity and selectivity (Figure 1b).

Prediction of absolute reactivity can be used to assess feasibility under process-relevant conditions,

while prediction of selectivity is key to reducing purification steps. Predictive tools therefore hold great

Page 3: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

2

promise for accelerating route and process development, ultimately delivering medicines to patients

both faster and at lower costs.

Figure 1. (a) Example of synthetic route to a drug compound. Prospects for AI-assisted route design. (b) Accurate prediction of reaction barriers gives both rate and selectivity. (c) The nucleophilic aromatic substitution (SNAr) reaction.

The current workhorse for prediction of organic reactions is density functional theory (DFT, Figure 2a).

Since rising to prominence in the early 90s, DFT has enjoyed extraordinary success in rationalizing

reactivity and selectivity across the reaction spectrum by modelling the full reaction mechanism.12 The

success of DFT can be traced in part due to a fortuitous cancellation of errors, which makes it particularly

suited for properties such as enantioselectivity, which depends on the relative energies of two

structurally very similar transition states (TSs). However, this cancellation of errors does not generally

extend to the prediction of the absolute magnitude of reactions barriers (activation free energies, ΔG‡).

In particular, DFT struggles with one very important class of reactions: ionic reactions in solution. Plata

and Singleton even suggested that computed mechanisms of this type can be so flawed that they are

“not even wrong”.13 Similarly, Maseras and co-workers only achieved agreement with experiment for

the simple condensation of an imine and an aldehyde in water by introducing an ad-hoc correction

factor, even when using more accurate methods than DFT.14 These results point to the fact that the

largest error in the DFT simulations is often due to the poor performance of the solvation model.

Page 4: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

3

Figure 2. Different types of quantitative reaction prediction approaches. Mechanistic DFT (a) and QSRR (b) are the current gold standard methods. Deep learning models (c) are emerging as an alternative. Hybrid models (d) combine mechanistic DFT modelling with traditional QSRR.

Machine learning represents a potential solution to the problems of DFT. Based on reaction data in

different solvents, machine learning models could in principle learn to compensate for both the

deficiencies in the DFT energies and the solvation model. Accurate QSRR machine learning models

(Figure 2b) have been constructed for, e.g., cycloaddition,15,16 SN2 substitution,17 and E2 elimination.18

While these models are highly encouraging, they treat simple reactions that occur in a single mechanistic

step and they are based on an amount of kinetic data (>500 samples) that is only available for very few

reaction classes. It is also not clear how they would incorporate the effects of more complex reaction

conditions, such as the use of catalysts or reagents. Another promising line of research uses machine

learning to predict DFT barrier heights and then use these barrier heights to predict experimental

outcomes. A recent study from Hong and co-workers used the ratio of predicted DFT barriers to predict

regioselectivity in radical C−H functionalization reactions.19 While these models can show good

performance, the predicted barriers still suffer from the shortcomings of the underlying DFT method and

solvation model. We therefore believe that for models to be broadly applicable in guiding experiments,

they should be trained to reproduce experimental rather than computed barrier heights.

Based on the recent success of machine learning for modelling reaction barriers, we wondered if we

could combine the traditional mechanistic modelling using DFT with machine learning in a hybrid method

(Figure 2d). Machine learning would here be used to correct for the deficiencies in the mechanistic

modelling. Hybrid models could potentially reach useful chemical accuracy (error below 1 kcal/mol)20,21

with fewer training data than QSRR models, be able to treat more complicated multi-step reactions, and

naturally incorporate the effect of catalysts directly in the DFT calculations. Mechanistic models are also

chemically understandable and the results can be presented to the chemist with both a view of the

computed mechanism and a value for the associated barrier. As a prototype application for a hybrid

model, we study the nucleophilic aromatic substitution (SNAr) reaction (Figure 1c), one of the most

important reactions in chemistry in general and the pharmaceutical chemistry in particular. The SNAr

reaction comprises 9% of all reaction carried out in pharma,22 and features heavily in commercial routes

to block-buster drugs.23,24 It has recently seen renewed academic interest concerning whether it occurs

through a stepwise or a concerted mechanism.25 We show that hybrid models for the SNAr reaction reach

chemical accuracy with ca 100-200 reactions in the training set, while traditional QSRR models based on

Page 5: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

4

quantum-chemical features seem to need at least 200 data points. Models based on purely structural

information such as reaction fingerprints need data in the range of 350–400 samples. Based on these

results, we envision a hierarchy of predictive models depending on how much data is available. Here,

transfer learning might ultimately represent the best of both worlds. Models pre-trained on a very large

number of DFT-calculated barriers26 can be retrained on a much smaller amount of high-quality

experimental data to achieve chemical accuracy for a wide range of reaction classes.

Results and Discussion First, we describe how the SNAr reaction dataset was collected and analysed. We then describe the

featurization of the reactions in terms of ground state and TS features. Machine learning models are

then built and validated. Finally, we use the model for regio- and chemoselectivity prediction on patent

reaction data, a task the model was not explicitly trained for.

Reaction dataset We collected 449 rate constants for SNAr reactions from the literature and ran 443 (98.7%) successfully

through the modelling procedure. Of these 443 reactions, 336 corresponded to unique reactions sets of

reactants and products, of which 274 were performed under only one set of conditions (temperature,

solvent) while 62 were performed under at least two sets of conditions. Activation energies were

obtained from the rate constants via the Eyring equation at the reaction temperature, and were in the

range 12.5–42.4 kcal/mol with a mean of 21.3 kcal/mol (Figure 3a). The dataset is diverse, with nitrogen,

oxygen and sulphur nucleophiles (Figure 3b) and oxygen, halogen and nitrogen leaving groups (Figure

3c), although the combinations of nucleophilic atom and leaving atoms is unevenly populated (Figure

3d). The two most common nucleophiles are piperidine (96 entries) and methoxide (49 entries), while

the most common substrates are dinitroarenes (Figure 3e). A principal components analysis of the full

feature space (vide infra) reveals a clear separation of reactions with respect to different nucleophile

and leaving group atom types (Figure 4). Supervised dimensionality reduction with partial least squares

(PLS) to 5 dimensions, followed by unsupervised dimensionality reduction with the uniform manifold

approximation and projection (UMAP) method27 yielded separated clusters with clear chemical

interpretation (Figure 4).

Page 6: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

5

Figure 3. Distribution of (a) activation energies (b) nucleophilic atoms and (c) leaving atoms in the dataset. (d) Number of reactions in the training set for combinations of nucleophilic atom and leaving atom (e) Frequency of the five most common nucleophiles and substrates in the training set.

Figure 4. Dataset visualization with unsupervised (PCA) and supervised (PLS + UMAP) dimensionality reduction for the Xfull feature set.

Page 7: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

6

Reaction feature generation To calculate the reaction features automatically, we constructed a workflow called predict-SNAr. The

workflow takes a reaction SMILES representation as input, deduces the reactive atoms and computes

reactants, transition states and products with a combination of semi-empirical methods and DFT (Figure

5a). Both concerted and stepwise mechanisms are treated and all structures are conformationally

sampled, making use of the lowest-energy conformer (Figure 5b). In the case of anionic nucleophiles, a

mixed explicit/implicit solvation model was employed to reduce the errors of the computed barriers.

The workflow also automatically calculates the quantum mechanical reaction features (Figure 5c). When

reactions corresponded to substitution on several electrophilic sites in a substrate, each reaction was

treated with a separate calculation.

Figure 5. (a) Automatic workflow for calculation of reaction mechanism and features (b) Important components of the workflow (c) Features calculated for ground and transition states calculated. Conformational sampling done with GFN2-xTB using the CREST tool, geometries refined with ωB97X-D/6-31+G(d) with SMD solvation. Final single points energies with ωB97X-D/6-311+G(d,p). Electronic structure features calculated with B3LYP/6-31+G(d).

For the hybrid model, we needed features for both the ground state molecules and the rate-determining

transition state. We opted for physical organic chemistry features which would be chemically

understandable and transferable to other reactions.28 We selected features associated with

nucleophilicity, electrophilicity, sterics, dispersion and bonding as well as features describing the solvent.

Page 8: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

7

As “hard” descriptors of nucleophilicity and electrophilicity, we used the surface average of the

electrostatic potential (V̅s) of the nucleophilic or electrophilic atom.29,30 (“Descriptor” and “feature” are

here used as synonyms.) As “soft” descriptors we used the atomic surface minimum of the average local

ionization energy (Is,min),31 as well as the local electron attachment energy (Es,min), which has been shown

to correlate well with SNAr reactivity.32 From conceptual DFT, we used the global electrophilicity

descriptor ω and nucleophilicity descriptor N, as well as the corresponding local nucleophilicity and

electrophilicity descriptors ℓN and ℓω.33 These electronic features were complemented with atomic

charges (q) from the DDEC6 scheme34 and the electrostatic potential at the nuclei (VN) of the reactive

atoms.35 In terms of sterics and dispersion, we use the ratio of solvent accessible surface area (SASAr)36

of the reactive atoms and the universal quantitative dispersion descriptor Pint.37 For bonding, we used

the DDEC6 bond orders (BO) of the carbon–nucleophile and carbon–leaving group bonds.38 Solvents

were described using the first five principal components (PC1–PC5) in the solvent database by Diorazio

and co-workers.39 The most important TS feature was the DFT-calculated activation free energy (ΔG‡DFT),

i.e., a “crude estimation” of the experimental target.40 We also added VN and DDEC6 charges and bond

orders at the TS geometry. We decided to not include the reaction temperature as one of the features,

as reactions with higher barriers tend to be run at higher temperatures as they are otherwise too slow,

and therefore correlate unduly with the target (see Supporting Information, section 5.6). Atomic

features are denoted with C (central), N (nucleophilic) or L (leaving) in parenthesis, e.g., q(C) for the

atomic charge of the central atom. Features at the TS geometry are indicated with a subscript “TS”, e.g.,

q(C)TS.

We chose to investigate three main feature sets: (1) Xfull, containing all the features (34 in total) for

maximum predictive accuracy, (2) XnoTS without any information from the TS, to assess whether hybrid

models are indeed more accurate, and (3) Xsmall, which represents a minimal set of 12 features that can

be interpreted more easily, with ΔG‡DFT as the only TS feature (see Supporting Information for complete

list). We also made two versions of XnoTS, excluding either “traditional” electronic descriptors (Xsurf,

missing V̅s, Is,min, Es,min and Pint) or surface features (Xtrad, missing ω, N, ℓω, ℓN, VN, q, and BO). As a

comparison to the physical organic features, we investigated four feature sets that only make use of the

2D structural information of the molecule: the Condensed Graph of Reaction (CGR) with the In Silico

Design and Data Analysis (ISIDA) descriptors41 (XISIDA,atom and XISIDA,seq), the Morgan reaction fingerprints42

as implemented in the RDKit (XMorgan),43 as well as the deep learning reaction fingerprints from Reymond

and co-workers (XBERT).44 These structural features can be calculated almost instantaneously and are

useful for fast prediction.

Choosing the best machine learning model We split the data into 80% used for model selection (training set) and 20% used to validate the final

model (test set). To compare the performance of a series of machine learning models on the training

set, we used bootstrap bias-corrected cross-validation (BBC-CV) with 10 folds.45 BBC-CV is an economical

alternative to nested cross-validation to avoid overfitting in the model selection process and also gives

an estimate of the bias for choosing the model that performs best on the training set. We measure

performance with the squared correlation coefficient (R2), the mean absolute error (MAE) and the root

mean squared error (RMSE). We focus on the MAE as its scale is directly comparable to the prediction

error. Error bars are given in terms of the standard errors.

Page 9: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

8

Figure 6. Model performance. (a) Selection of hybrid models of increasing complexity. (b) Performance of DFT versus hybrid model. (c) Comparison of physical organic feature sets. (d) Comparison of structural feature sets. Error bar corresponds to one standard error. Abbreviations explained in the text.

The results of the model validation (Figure 6a, Table S5) show a clear progression from simpler models

such as linear regression (LR) with a MAE of 1.20 kcal/mol, to intermediate methods such as random

forests (RF) at 0.98 kcal/mol, to more advanced Support Vector Regression (SVR) and Gaussian Process

Regression (GPR) models at 0.80 kcal/mol. Most importantly, the best methods are well below chemical

accuracy (1 kcal/mol). In comparison, the raw DFT barriers ΔG‡DFT show a high MAE of 2.93 kcal/mol and

have the same predictive value as just guessing the mean of the training dataset (Figure 6b).

Compensating for systematic errors in the DFT energies by linear correction helps, but still has an

unacceptable MAE of 1.74 kcal/mol. Interestingly, simpler linear methods such as the Automatic

Relevance Determination (ARD) can achieve the same performance as the non-linear RF when

polynomial and interaction features of second order (PF2) are used to capture non-linear effects. The

overall best method considering MAE is GPR with the Matern 3/2 kernel (GPRM3/2). This result is very

gratifying as GPRs are resistant to overfitting, do hyperparameter tuning internally and produce error

bars that can be used for risk assessment. We therefore selected GPRM3/2 as our final method and also

used it to make comparisons between different feature sets. In the BBC-CV evaluation, it had an R2 of

0.87, an MAE of 0.80 kcal/mol, and an RMSE of 1.41 kcal/mol. Importantly, the BBC-CV method indicates

a very low bias of only 0.02 kcal/mol for MAE in choosing GPRM3/2 as the best method, so overfitting in

the model selection can be expected to be small. Indeed, GPRM3/2 shows excellent performance on the

external test set, with a R2 of 0.93, a MAE of 0.77 kcal/mol and an RMSE of 1.01 kcal/mol (Figure 7). The

prediction intervals measuring the uncertainty of the individual prediction have a coverage of 99% for

the test set, showing that the model can also accurately assess how reliable its predictions are.

Page 10: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

9

Figure 7. Performance of final GPRM3/2 model on the external test set. Coverage for test set: 99%.

Performance of different feature sets Now, are hybrid models using TS features better than the traditional QSRR models based on just ground-

state features? The validation results indicate that hybrid models built with the full feature set Xfull

perform the best, but not significantly better than models built on XNoTS without TS features (Figure 6c).

Also the model built on the small expert-chosen feature set Xsmall shows similar performance. To

investigate the matter more deeply, we calculated the learning curves of GPRM3/2 using the different

features sets (Figure 8a) on the full dataset. We see that all models indeed do perform similarly with the

amount of training data used for the model selection (318 reactions), but that the hybrid models based

on Xfull and Xsmall seem to have an advantage below 150 samples. Indeed, the learning curve for predicting

the ΔG‡DFT based on XNoTS also starts levelling off after ca 150 samples (Figure 8b). This indicates that the

model is able to implicitly learn ΔG‡DFT from the ground state features given sufficient data. It seems that

there is still a residual advantage using the full hybrid model even with larger dataset sizes, although this

advantage becomes smaller and smaller. Models built on ground state features lacking either surface

features (Xtrad) or lacking more traditional electronic descriptors (Xsurf) showed worse performance (MAE:

1.00 and 1.10 kcal/mol, respectively) than when both were included as in XnoTS (MAE: 0.86 kcal/mol).

Therefore, both should be included for maximum performance and seem to capture different aspects of

reactivity.

How good can the model get given even more data? Any machine learning model is limited by the

intrinsic noise of the underlying training data, given in our case by the experimental error of the kinetic

measurement. In the dataset, there are four reactions reported with the same solvent and temperature

but in different labs or on different occasions. Differences between the activation energies are 0.1, 0.1,

0.5 and 1.6 kcal/mol. The larger difference is probably an outlier, and we estimate the experimental

error is on the order 0.1–0.5 kcal/mol. In comparison, the interlaboratory error for a data set of SN2

reactions was estimated to ca 0.7 kcal/mol.17 It is thus reasonable to believe that the current model with

a MAE of 0.77 kcal/mol on the external test set is getting close to the performance that can be achieved

given the quality of the underlying data. Gathering more data is therefore not expected to significantly

improve the accuracy of the average prediction, but may widen the applicability domain by covering a

broader range of structures (vide infra) and reduce the number of outlier predictions.

Page 11: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

10

Figure 8. (a) Learning curve giving the mean absolute error as a function of number of reactions in the training set. (b) Learning curve to predict the DFT activation energies using the ground state features.

Models based on structural information Given the time needed to develop both traditional and hybrid QSRR models, an attractive option is using

features derived from just chemical connectivity. We investigated this option using the CGR/ISIDA

approach with either atom-centred (XISIDA,atom) or sequential (XISIDA,seq) fragment features, as well as

reaction difference fingerprints of the Morgan type with a radius of three atoms (XMorgan3). The results

with GPRM3/2 show good performance using XMorgan3, almost reaching chemical accuracy with a MAE of

1.09 kcal/mol (Figure 6d). ISIDA sequence and atom features perform worse (Figure 6d). They also

require an accurate atom-mapping of the reaction, and we found that automatic atom mapping failed

for 50 (11%) of the reactions studied. The methods above rely on expert-crafted algorithms to generate

fingerprints or feature vectors. In recent years, deep learning has emerged as a method for creating such

representations from the data itself, i.e., representation learning. The recent reaction fingerprint from

Reymond and co-workers is one such example, where the fingerprint is learned by a BERT deep learning

model that is pre-trained in an unsupervised manner on reaction SMILES from patent data.44 Gratifyingly,

the model built on the XBERT feature set performs on par or slightly better than XMorgan3 with an MAE of

1.03 kcal/mol (Figure 6d). The big advantage with the BERT fingerprint is that it does not need atom

mapping and can potentially be used with noisy reaction SMILES with no clear separation between

reactants, reagents, catalysts and solvents. We also used BERT fingerprints specially tuned for reaction

classification to construct reaction maps, which show clear separation between different types of

nucleophiles and electrophiles in the dataset (Figure S6). The learning curve shows that the model based

on XBERT is more data-hungry than the models based on physical organic features, and requires ca 350–

400 data points to reach chemical accuracy (Figure 8a). The radius of three for the Morgan fingerprint

was chosen to incorporate long-range effects of electron-donating and electron-withdrawing groups in

the para position of the aromatic ring (Figure 9b). Plotting the MAE as a function of the fingerprint radius

shows a clear minimum at a radius of three (Figure 9a). With this radius, all relevant reaction information

seems to be captured, and increasing the radius further probably just adds noise through bit clashes in

the fingerprint generation. Encouraged by the promising results with the structural data, we wondered

if a combined feature set with both the physical organic features in Xfull and the structural features in

XBERT would perform even better. However, the model built on the combined feature set performs on

par with the model built on only Xfull (Table S5)

In summary, models based on reaction fingerprints are an attractive alternative when a sizeable dataset

of at least 350 reactions are available as they are easy to develop and make very fast predictions.

Page 12: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

11

Figure 9. (a) Mean absolute error for GPRM3/2 as a function of Morgan fingerprint radius. (b) Atoms and bonds considered by Morgan fingerprints of different radius. A radius of radius three or more is needed to capture effect from groups in the para position of the aromatic ring.

Interpretability There has a been a push in the machine learning community in recent years to not only predict accurately

but also to understand the factors behind the prediction.46 Models are often interpreted in terms of their

feature importances, i.e., how much a certain feature contributes to the prediction. Feature importances

can be obtained directly from multivariate linear regression models as the regression coefficients and

have been used extensively to give insight on reaction mechanisms based on ground-state features.47 A

number of modern techniques can obtain feature importances for any machine learning technique,

including SHAP values48 and permutation importances,49 potentially allowing the simple interpretation

of linear models to be combined with the higher accuracy of more modern non-linear methods.

Although feature importances can be easily calculated, they are not always easily interpretable. In

particular, correlation between features poses severe problems. This problem is present for our feature

set, as shown by the Spearman rank correlation matrix and the variance inflation factors (VIFs) of Xfull

(Figure S7). Therefore, special care has to be taken when calculating the feature importances, and the

final interpretation of them will not be straightforward. To get around this technical problem of

multicollinearity, we clustered the features based on their Spearman rank correlations, keeping only one

of the features in each cluster (Figure 10a). The stability of the feature ranking was further analysed by

using ten bootstrap samples of the data.

Page 13: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

12

Figure 10. (a) Clustering of correlated features based on Spearman rank correlation (b) Bootstrapped feature ranking for clustered Xfull and XNoTS. For identity of feature clusters, see the Supporting Information.

First, we looked at how important ΔG‡DFT is in our hybrid model. It turns out that it is consistently ranked

as the most important feature across all bootstrap samples (Figure 10b). The second-most important

feature is the soft electrophilicity feature Es,min(C). To see more clearly which are the most important

ground-state features, we analysed a model built on XnoTS (Figure 10b). The most important feature is

again the soft electrophilicity descriptor Es,min(C). Other important features are the soft nucleophilicity

descriptor Is,min(N) and the feature cluster of hard electrophilicity represented by V̅s(C). The global

nucleophilicity descriptor N and the electrophile-nucleophile bond strength through BO(C–N) are also

ranked consistently high.

In summary, the most important feature for the hybrid models is, as expected, the DFT-computed

activation free energy ΔG‡DFT. Models trained without the TS features give insight into the features of

substrates and nucleophiles that govern reactivity. Here, the most important features are the

electrophilicity of the central carbon atom of the substrate, followed by those related to nucleophilicity.

Steric and solvent features are less important. A plausible reason that solvent features have lower

importance is that the effect of the solvent is already incorporated in the ΔG‡DFT as well as in the influence

of the implicit solvent models on the other quantum-mechanically derived features. Steric features may

become more important for other types of substrates.

Applicability domain The applicability domain (AD) is a central concept to any QSRR model used for regulatory purposes

according to the OECD guidelines.50 The AD is broadly defined by the OECD as “the response and chemical

structure space in which the model makes predictions with a given reliability”. Although there have been

many attempts to define the AD for prediction of molecular properties, there has been little work for

reaction prediction models. Here, we will follow a practical approach to (1) define a set of strict rules for

when the model shouldn’t be applied based on the reaction type and the identity of the reactive atoms,

(2) identify potentially problematic structural motifs from visualization of outliers, and (3) assess

whether the uncertainty provided by the GPRM3/2 model can be used for risk assessment.

Page 14: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

13

First, the model should only be applied to SNAr reactions. Second, the model can only be used with

confidence for those reactive atom types that are part of the training set (Figure 3b). One example of

reactions falling outside the applicability domain according to these rules are those with anionic carbon

nucleophiles. A visual inspection of the outliers (residual of > 2 kcal/mol) for the training set identifies

many reactions involving methoxide nucleophile. This tendency can be understood based on the poor

performance of implicit solvation models for such small, anionic nucleophiles. Some of these reactions

are also very slow and have been run at high temperature and the rate constants have been determined

through extrapolation techniques, leading to a larger error in the corresponding rates. Outliers from the

test set involve azide and secondary amine nucleophiles. However, secondary amines are also the most

common type of nucleophile in the dataset and are therefore expected to contribute to some outliers.

For a complete list of all outliers, see the Supporting Information.

Figure 11. Example of outliers for training and test set. Ranges in parenthesis correspond to errors in kcal/mol.

We also experimented with using a similarity measures to decide whether the prediction for a certain

reaction should be trusted or not. For molecular properties, similarity to the training set is usually

measured by distance metrics based on the molecular fingerprints.51 For reactions, difference

fingerprints have been shown to differentiate between reaction classes, but their suitability for defining

an applicability domain within one reaction class is not clear.52 In the absence of a clear similarity metric,

we used the Distance to Model in X-space (DModX) of a PLS model with two components. We then

compared the performance of DModX to the standard deviation (std) of the GPRM3/2 predictions using

integral accuracy curves (Figure 12). In these curves, the predicted values are ordered from most reliable

to least reliable based on the uncertainty metric. The MAE is then plotted as a function of the portion of

included data. A good uncertainty metric should show a curve with an upward slope from left to right,

as including more points with larger uncertainty should lead to a larger error. As can be seen from the

plot, both DModX and GPRM3/2 std are decent measures of uncertainty. As the GPRM3/2 std performs best,

the prediction intervals (as shown in Figure 7) can be used directly as a measure of how reliable the

prediction is.

Page 15: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

14

Figure 12. Integral accuracy curve. Predicted values are ordered according to an uncertainty measure from most certain to least certain. MAE is calculated from the predictions of the GPRM3/2 model.

As there are 62 reactions which occur in the dataset with more than one reaction condition (different

temperature or solvent), we investigated the performance of leaving these reactions out altogether in

the modelling. With this leave-one-reaction-out validation approach, we observed a MAE of 1.00

kcal/mol for GPRM3/2 (compared to 0.80 kcal/mol from normal cross-validation). We also tested leave-

one-substrate-out, giving a MAE of 1.20 kcal/mol and leave-one-nucleophile-out, giving a of MAE 0.68

kcal/mol. These results indicate that the model is able to predict outside its immediate chemical space

with good accuracy, and not only interpolate.

Taken together, the applicability domain of our model is defined strictly in terms of the type of reaction

(SNAr) and the types of reactive atoms in the training set. The outlier analysis identified that extra care

should be taken when interpreting the results of reactions at high temperature and with certain

nucleophile classes. The width of the prediction interval from the GPRM3/2 model is a useful measure of

the uncertainty of the prediction.

Validation on selectivity data To assess whether the model can be used not only for predicting rates, but also selectivities, we compiled

a dataset of reactions with potential regio- or chemoselectivity issues from patent data. A total of 4365

SNAr reactions were considered, of which 1214 had different reactive sites on the same aromatic ring,

while only one product was recorded experimentally. A reactive site was considered as a ring carbon

with a halogen subsituent (F, Cl, Br and I) that could be substituted with SNAr. Regioselectivity means

distinguishing between reactive sites with the same type of halogen, while for chemoselectivity the

halogens are different. Out of these 1214, we selected the 100 with lowest product molecular weight

for a preliminary evaluation. As reaction solvent or temperature was not available, we used acetonitrile

for neutral reactions and methanol for ionic reactions and a reaction temperature of 25 °C. Each possible

reaction leading to different isomers was modelled by a separate predict-SNAr calculation (209 in total),

and the predicted major isomer was taken as the one corresponding to the lowest ΔG‡. As another

testament to the robustness of the workflow, only 3 of 209 TS calculations failed (1.4%), leading finally

to results for 97 reactions with selectivity issues. Of these, 66 involved regioselectivity and 31

chemoselectivity. The results show that GPRM3/2 was able to predict the correct site in 86% of the cases

(top-1 accuracy), with 85% for regioselectivity and 87% for chemoselectivity. In comparison, predictions

based on only the DFT activation energies give a comparable top-1 accuracy of 87%, with 91% for

regioselectivity and 77% for chemoselectivity. It is notable that the GPRM3/2 model shows a much better

score for chemoselectivity (87%) than DFT (77%). For regioselectivity prediction, where the same

element is being substituted, DFT probably works well due to error cancellation. That the hybrid ML

model performs better than DFT for chemoselectivity clearly shows that it has learnt to compensate for

Page 16: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

15

the systematic errors in the DFT calculations. In fact, it seems that the ML model is balancing the gain in

chemoselectivity prediction with a loss in regioselectivity (85%) compared to DFT (91%), leading to a

comparable accuracy on both tasks. More training data is expected to alleviate this loss in regioselectivity

accuracy. Also, inclusion of the actual solvents and temperatures could also increase accuracy.

Overall, it is remarkable that the hybrid model GPRM3/2 performs so well (86%) for selectivity as it was

not explicitly trained on this task. In comparison, for electrophilic aromatic substitution (SEAr), single-

task deep learning models trained for selectivity prediction of bromination, chlorination, nitration and

sulfonylation achieved top-1 accuracies of 50–87%.53 For bromination, the RegioSQM model based on

semiempirical energies of the regioisomeric intermediate σ-complexes achieved 80%, compared to 85%

for the neural network just mentioned. In light of these models, the 86% top-1 accuracy obtained with

GPRM3/2 for SNAr looks very competitive. Likely, hybrid models explicitly trained for selectivity prediction

can perform even better. Another approach would be to use transfer learning to repurpose deep

learning models for barrier prediction to selectivity prediction.

Conclusions and outlook We have created hybrid mechanistic/machine learning models for the SNAr reaction which incorporate

TS features in addition to the traditional physical organic features of reactants and products. The chosen

Gaussian Process Regression model achieved a mean absolute error of only 0.77 kcal/mol on the external

test set, well below the targeted chemical accuracy of 1 kcal/mol. Furthermore, the model achieves a

top-1 accuracy for regio- and chemoselectivity prediction on patent reaction data of 86%, without being

explicitly trained for this task. Finally, the model comes with a clear applicability domain specification

and prediction error bars that enables the end user to make a contextualized risk assessment depending

on what accuracy is required.

By studying models built on reduced sets of the physical organic features, as well as reaction fingerprints,

we identified separate data regimes for modelling the SNAr reaction (Figure 13). In the range 0–50

samples, it is questionable whether accurate and generalizable machine learning models can be

constructed.54 Instead, we suggest that traditional mechanistic modelling with DFT should be used, with

appropriate consideration of their weaknesses. With 50–150 samples, hybrid models are likely the most

accurate choice, and should be used if time and resources for their development is available. In the range

150–300 samples, traditional QSRR models based on physical organic features reach similar accuracy as

hybrid models, while models based on purely structural information become competitive with over 300

samples. It will be very interesting to see if these numbers generalize to other reaction classes. For

choosing features for QSRR models, it is notable that electronic reactivity features from DFT were

consistently ranked high in the feature importances, and we saw that the inclusion of surface features

makes the models significantly better.

Figure 13. Models for quantitative rate prediction for different data regimes based on the SNAr dataset.

Our workflow can handle the mechanistic spectrum of concerted and stepwise SNAr reactions, and we

are currently working on extending it to handle the influence of general or specific acid and base catalysts

as well as treating related reaction classes. This general model is a significant improvement in generality

compared to previous work, which modelled selectivity of SNAr reactions in terms of the relative stability

of the σ addition complexes and therefore was only applicable to reactions with step-wise mechanisms. 55–57 Although promising, we believe that widespread use of hybrid models is currently held back by

Page 17: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

16

difficulties in computing transition states in an effective and reliable way. We envision that this problem

will be solved in the near future by deep learning approaches that can predict both TS geometries58 and

DFT-computed barriers26 based on large, publicly available datasets.59,60 In the end, machine learning for

reaction prediction needs to reproduce experiment, and transfer learning will probably be key to utilizing

small high-quality kinetic datasets together with large amounts of computationally generated data.

Regardless of their construction, accurate reaction prediction models will be key components of

accelerated route design, reaction optimization and process design enabling the delivery of medicines

to patients faster and with reduced costs.

Acknowledgements Stefan Grimme is kindly acknowledged for permission to use the xtb code and for technical advice

regarding its use. Ramil Nugmanov is acknowledged for advice on calculation of the ISIDA descriptors.

Tore Brinck acknowledges support from the Swedish Research Council (VR).

Supporting information Detailed computational methods, full description of machine learning workflow and results. Description

of the experimental and computed datasets.

References 1. Muratov, E.N., Bajorath, J., Sheridan, R.P., Tetko, I.V., Filimonov, D., Poroikov, V., Oprea, T.I.,

Baskin, I.I., Varnek, A., Roitberg, A., et al. (2020). QSAR without borders. Chem. Soc. Rev. 49, 3525–3564.

2. Coley, C.W., Eyke, N.S., and Jensen, K.F. (2019). Autonomous discovery in the chemical sciences

part II: Outlook. Angew. Chem. Int. Ed.

3. Engkvist, O., Norrby, P.-O., Selmi, N., Lam, Y., Peng, Z., Sherer, E.C., Amberg, W., Erhard, T., and

Smyth, L.A. (2018). Computational prediction of chemical reactions: current status and outlook. Drug

Discov. Today 23, 1203–1218.

4. Ahneman, D.T., Estrada, J.G., Lin, S., Dreher, S.D., and Doyle, A.G. (2018). Predicting reaction

performance in C–N cross-coupling using machine learning. Science 360, 186–190.

5. Sandfort, F., Strieth-Kalthoff, F., Kühnemund, M., Beecks, C., and Glorius, F. A Structure-Based

Platform for Predicting Chemical Reactivity. Chem.

6. Zahrt, A.F., Henle, J.J., Rose, B.T., Wang, Y., Darrow, W.T., and Denmark, S.E. (2019). Prediction

of higher-selectivity catalysts by computer-driven workflow and machine learning. Science 363,

eaau5631.

7. Reid, J.P., and Sigman, M.S. (2019). Holistic prediction of enantioselectivity in asymmetric

catalysis. Nature 571, 343–348.

8. Schwaller, P., Laino, T., Gaudin, T., Bolgar, P., Hunter, C.A., Bekas, C., and Lee, A.A. (2019).

Molecular Transformer: A Model for Uncertainty-Calibrated Chemical Reaction Prediction. ACS Cent. Sci.

5, 1572–1583.

9. Klucznik, T., Mikulak-Klucznik, B., McCormack, M.P., Lima, H., Szymkuć, S., Bhowmick, M., Molga,

K., Zhou, Y., Rickershauser, L., Gajewska, E.P., et al. (2018). Efficient Syntheses of Diverse, Medicinally

Relevant Targets Planned by Computer and Executed in the Laboratory. Chem 4, 522–532.

10. Tomberg, A., Johansson, M.J., and Norrby, P.-O. (2019). A Predictive Tool for Electrophilic

Aromatic Substitutions Using Machine Learning. J. Org. Chem. 84, 4695–4703.

11. Tomberg, A., Muratore, M.É., Johansson, M.J., Terstiege, I., Sköld, C., and Norrby, P.-O. (2019).

Relative Strength of Common Directing Groups in Palladium-Catalyzed Aromatic C−H Activation. iScience

20, 373–391.

Page 18: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

17

12. Vogel, P., and Houk, K.N. (2019). Organic Chemistry: Theory, Reactivity and Mechanisms in

Modern Synthesis (Wiley).

13. Plata, R.E., and Singleton, D.A. (2015). A Case Study of the Mechanism of Alcohol-Mediated

Morita Baylis–Hillman Reactions. The Importance of Experimental Observations. J. Am. Chem. Soc. 137,

3811–3826.

14. Pérez-Soto, R., Besora, M., and Maseras, F. (2020). The Challenge of Reproducing with

Calculations Raw Experimental Kinetic Data for an Organic Reaction. Org. Lett. 22, 2873–2877.

15. Ravasco, J.M.J.M., and Coelho, J.A.S. (2020). Predictive Multivariate Models for Bioorthogonal

Inverse-Electron Demand Diels–Alder Reactions. J. Am. Chem. Soc. 142, 4235–4241.

16. Glavatskikh, M., Madzhidov, T., Horvath, D., Nugmanov, R., Gimadiev, T., Malakhova, D.,

Marcou, G., and Varnek, A. (2019). Predictive Models for Kinetic Parameters of Cycloaddition Reactions.

Mol. Inform. 38, 1800077.

17. Gimadiev, T., Madzhidov, T., Tetko, I., Nugmanov, R., Casciuc, I., Klimchuk, O., Bodrov, A.,

Polishchuk, P., Antipin, I., and Varnek, A. (2019). Bimolecular Nucleophilic Substitution Reactions:

Predictive Models for Rate Constants and Molecular Reaction Pairs Analysis. Mol. Inform. 38, 1800104.

18. Madzhidov, T.I., Bodrov, A.V., Gimadiev, T.R., Nugmanov, R.I., Antipin, I.S., and Varnek, A.A.

(2015). Structure–reactivity relationship in bimolecular elimination reactions based on the condensed

graph of a reaction. J. Struct. Chem. 56, 1227–1234.

19. Li, X., Zhang, S.-Q., Xu, L.-C., and Hong, X. (2020). Predicting Regioselectivity in Radical C−H

Functionalization of Heterocycles through Machine Learning. Angew. Chem. Int. Ed.

20. Houk, K.N., and Liu, F. (2017). Holy Grails for Computational Organic Chemistry and

Biochemistry. Acc. Chem. Res. 50, 539–543.

21. Peterson, K.A., Feller, D., and Dixon, D.A. (2012). Chemical accuracy in ab initio thermochemistry

and spectroscopy: current strategies and future challenges. Theor. Chem. Acc. 131, 1079.

22. Boström, J., Brown, D.G., Young, R.J., and Keserü, G.M. (2018). Expanding the medicinal

chemistry synthetic toolbox. Nat. Rev. Drug Discov. 17, 709–727.

23. Finlay, M.R.V., Anderton, M., Ashton, S., Ballard, P., Bethel, P.A., Box, M.R., Bradbury, R.H.,

Brown, S.J., Butterworth, S., Campbell, A., et al. (2014). Discovery of a Potent and Selective EGFR Inhibitor

(AZD9291) of Both Sensitizing and T790M Resistance Mutations That Spares the Wild Type Form of the

Receptor. J. Med. Chem. 57, 8249–8267.

24. Baumann, M., and Baxendale, I.R. (2013). An overview of the synthetic routes to the best selling

drugs containing 6-membered heterocycles. Beilstein J. Org. Chem. 9, 2265–2319.

25. Kwan, E.E., Zeng, Y., Besser, H.A., and Jacobsen, E.N. (2018). Concerted nucleophilic aromatic

substitutions. Nat. Chem. 10, 917–923.

26. Grambow, C.A., Pattanaik, L., and Green, W.H. (2020). Deep Learning of Activation Energies. J.

Phys. Chem. Lett. 11, 2992–2997.

27. McInnes, L., Healy, J., and Melville, J. (2018). UMAP: Uniform Manifold Approximation and

Projection for Dimension Reduction. arXiv:1802.03426.

28. Beker, W., Gajewska, E.P., Badowski, T., and Grzybowski, B.A. (2019). Prediction of Major Regio‐

, Site‐, and Diastereoisomers in Diels–Alder Reactions by Using Machine‐Learning: The Importance of

Physically Meaningful Descriptors. Angew. Chem. Int. Ed. 58, 4515–4519.

29. Murray, J.S., and Politzer, P. (2011). The electrostatic potential: an overview. Wiley Interdiscip.

Rev. Comput. Mol. Sci. 1, 153–163.

30. Brinck, T., Jin, P., Ma, Y., Murray, J.S., and Politzer, P. (2003). Segmental analysis of molecular

surface electrostatic potentials: application to enzyme inhibition. J. Mol. Model. 9, 77–83.

31. Sjoberg, P., Murray, J.S., Brinck, T., and Politzer, P. (1990). Average local ionization energies on

the molecular surfaces of aromatic systems as guides to chemical reactivity. Can. J. Chem. 68, 1440–

1443.

Page 19: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

18

32. Brinck, T., Carlqvist, P., and Stenlid, J.H. (2016). Local Electron Attachment Energy and Its Use for

Predicting Nucleophilic Reactions and Halogen Bonding. J. Phys. Chem. A 120, 10023–10032.

33. Oller, J., Pérez, P., Ayers, P.W., and Vöhringer-Martinez, E. (2018). Global and local reactivity

descriptors based on quadratic and linear energy models for α,β-unsaturated organic compounds. Int. J.

Quantum Chem. 118, e25706.

34. Manz, T.A., and Gabaldon Limas, N. (2016). Introducing DDEC6 atomic population analysis: part

1. Charge partitioning theory and methodology. RSC Adv. 6, 47771–47801.

35. Galabov, B., Ilieva, S., Koleva, G., Allen, W.D., Schaefer III, H.F., and Schleyer, P. von R. (2013).

Structure–reactivity relationships for aromatic molecules: electrostatic potentials at nuclei and

electrophile affinity indices. Wiley Interdiscip. Rev. Comput. Mol. Sci. 3, 37–55.

36. Lee, B., and Richards, F.M. (1971). The interpretation of protein structures: Estimation of static

accessibility. J. Mol. Biol. 55, 379-IN4.

37. Pollice, R., and Chen, P. (2019). A Universal Quantitative Descriptor of the Dispersion Interaction

Potential. Angew. Chem. Int. Ed. 58, 9758–9769.

38. Manz, T.A. (2017). Introducing DDEC6 atomic population analysis: part 3. Comprehensive

method to compute bond orders. RSC Adv. 7, 45552–45581.

39. Diorazio, L.J., Hose, D.R.J., and Adlington, N.K. (2016). Toward a More Holistic Framework for

Solvent Selection. Org. Process Res. Dev. 20, 760–773.

40. Zhang, Y., and Ling, C. (2018). A strategy to apply machine learning to small datasets in materials

science. Npj Comput. Mater. 4, 25.

41. Varnek, A., Fourches, D., Hoonakker, F., and Solov’ev, V.P. (2005). Substructural fragments: an

universal language to encode reactions, molecular and supramolecular structures. J. Comput. Aided Mol.

Des. 19, 693–703.

42. Rogers, D., and Hahn, M. (2010). Extended-Connectivity Fingerprints. J. Chem. Inf. Model. 50,

742–754.

43. RDKit: Open-source cheminformatics http://www.rdkit.org.

44. Philippe Schwaller, Daniel Probst, Alain C. Vaucher, Vishnu H Nair, Teodoro Laino, and Jean-Louis

Reymond (2019). Data-Driven Chemical Reaction Classification, Fingerprinting and Clustering using

Attention-Based Neural Networks. ChemRxiv.

45. Tsamardinos, I., Greasidou, E., and Borboudakis, G. (2018). Bootstrapping the out-of-sample

predictions for efficient and accurate cross-validation. Mach. Learn. 107, 1895–1922.

46. Molnar, C. (2019). Interpretable Machine Learning: A Guide for Making Black Box Models

Explainable.

47. Sigman, M.S., Harper, K.C., Bess, E.N., and Milo, A. (2016). The Development of Multidimensional

Analysis Tools for Asymmetric Catalysis and Beyond. Acc. Chem. Res. 49, 1292–1301.

48. Lundberg, S.M., and Lee, S.-I. (2017). A Unified Approach to Interpreting Model Predictions. In

Advances in Neural Information Processing Systems 30, I. Guyon, U. V. Luxburg, S. Bengio, H. Wallach, R.

Fergus, S. Vishwanathan, and R. Garnett, eds. (Curran Associates, Inc.), pp. 4765–4774.

49. Breiman, L. (2001). Random Forests. Mach. Learn. 45, 5–32.

50. OECD (2014). Guidance Document on the Validation of (Quantitative) Structure-Activity

Relationship [(Q)SAR] Models.

51. Hanser, T., Barber, C., Marchaland, J.F., and Werner, S. (2016). Applicability domain: towards a

more formal definition. SAR QSAR Environ. Res. 27, 865–881.

52. Schneider, N., Lowe, D.M., Sayle, R.A., and Landrum, G.A. (2015). Development of a Novel

Fingerprint for Chemical Reactions and Its Application to Large-Scale Reaction Classification and

Similarity. J. Chem. Inf. Model. 55, 39–53.

53. Struble, T.J., Coley, C.W., and Jensen, K.F. (2020). Multitask prediction of site selectivity in

aromatic C–H functionalization reactions. React. Chem. Eng. 5, 896–902.

Page 20: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

19

54. Yaser S. Abu-Mostafa, Magdon-Ismail, M., and Lin, H.T. (2012). Learning from Data: A Short

Course (AMLBook.com).

55. Liljenberg, M., Brinck, T., Herschend, B., Rein, T., Rockwell, G., and Svensson, M. (2011). A

pragmatic procedure for predicting regioselectivity in nucleophilic substitution of aromatic fluorides.

Tetrahedron Lett. 52, 3150–3153.

56. Liljenberg, M., Brinck, T., Herschend, B., Rein, T., Tomasi, S., and Svensson, M. (2012). Predicting

Regioselectivity in Nucleophilic Aromatic Substitution. J. Org. Chem. 77, 3262–3269.

57. Liljenberg, M., Brinck, T., Rein, T., and Svensson, M. (2013). Utilizing the σ-complex stability for

quantifying reactivity in nucleophilic substitution of aromatic fluorides. Beilstein J. Org. Chem. 9, 791–

799.

58. Pattanaik, L., Ingraham, J., Grambow, C., and Green, W.H. (2020). Generating Transition States

of Isomerization Reactions with Deep Learning. ChemRxiv.

59. Grambow, C.A., Pattanaik, L., and Green, W.H. (2020). Reactants, products, and transition states

of elementary chemical reactions based on quantum chemistry. Sci. Data 7, 137.

60. von Rudorff, G.F., Heinen, S.N., Bragato, M., and von Lilienfeld, O.A. (2020). Thousands of

reactants and transition states for competing E2 and SN2 reactions.

Page 22: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

1

Supporting Information

1 Contents 1 Contents .............................................................................................................................................. 1

2 Predict-SNAr workflow ........................................................................................................................ 2

2.1 Input preparation ........................................................................................................................ 2

2.2 Input parsing and structure generation ...................................................................................... 3

2.3 Mechanistic calculations ............................................................................................................. 3

2.3.1 Quantum chemistry ............................................................................................................ 3

2.3.2 Transition state calculations ............................................................................................... 4

2.3.3 Use of additional constraints for xtb optimizations............................................................ 5

2.3.4 Calculation monitor ............................................................................................................ 6

2.3.5 Explicit solvation ................................................................................................................. 6

2.3.6 Use of electronic temperature in xtb calculations .............................................................. 8

2.4 Feature generation ..................................................................................................................... 8

2.5 Running the workflow ............................................................................................................... 10

3 Machine learning .............................................................................................................................. 11

3.1 Model selection ........................................................................................................................ 11

3.1.1 Models tested ................................................................................................................... 14

3.2 Dataset visualization ................................................................................................................. 14

3.3 Feature importances ................................................................................................................. 15

3.4 Learning curves ......................................................................................................................... 20

3.5 Analysis of reactions with large errors ..................................................................................... 20

3.6 Dependence on number of heavy atoms ................................................................................. 22

3.7 Y-randomization ........................................................................................................................ 22

4 Regio- and chemoselectivity validation ............................................................................................ 22

5 Experimental dataset ........................................................................................................................ 24

5.1 Description of the modelling data ............................................................................................ 24

5.2 Distribution of activation free energies .................................................................................... 25

5.3 Most common substrates and nucleophiles ............................................................................. 26

5.4 Replicated reactions ................................................................................................................. 27

5.5 Conditions ................................................................................................................................. 27

5.6 Reaction temperatures ............................................................................................................. 28

6 Workflow database ........................................................................................................................... 29

7 Analysis package ............................................................................................................................... 31

8 References ........................................................................................................................................ 31

Page 23: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

2

2 Predict-SNAr workflow The predict-SNAr workflow takes a reaction SMILES1 as input, identifies the nucleophile, substrate,

product and leaving group and then sets up and performs all calculations of the mechanism and the

descriptors (Figure S1). Temperature and solvent are also used throughout the calculations. The results

are stored in a database for easy retrieval. predict-SNAr has been optimized for robustness and includes

extensive error checks for the quantum chemical calculations. Below, a detailed description of the

different steps will be given.

Figure S1. Overview of predict-SNAr workflow

2.1 Input preparation The literature data from the file “kinetic_data_v4.xlsx” was first pruned to remove entries for which

either activation free energy, solvent or temperature was missing. (After the modelling was completed,

some additional entries have been added in database, see section 5.)

Solvent mixtures were processed to determine the “influential solvent” as the SMD solvent model (vide

infra) can only handle single solvents. It is clear that solvent properties are not a simple linear

interpolation between the properties of the constituent solvents. Determining which single solvent to

substitute for a solvent mixture is somewhat arbitrary, but we used two principles to guide our

reasoning: (1) preferential solvation and (2) activity.2 Preferential solvation means that ions will be

preferentially solvated by the solvent to which they have the strongest interactions. More polar solvents

should therefore have a larger influence in solvating ionic reactants than expected based on their molar

fraction in comparison with less polar solvents. Activity coefficients will be higher for minority solvents,

meaning that they will exert a higher “effective” mol fraction than the raw numbers indicate. By

combining these two principles, we came up with the following rule of thumb for binary solvent

mixtures: if the polar solvent has a mole fraction of at least 0.2, it will be used as the single solvent in the

workflow, otherwise the less polar solvent will be used.

Input

•Reaction SMILES

•Temperature

•Solvent

Pre-processing

•Identify reactive molecules

•Identify reactive atoms

•Generate 3D structures

Mechanistic calculations

•Find and optimize intermediates and transition states

•Optimize reactants and products

•Explicit solvation calculations

Descriptor calculation

Output

•Write to workflow database

Page 24: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

3

The input preparation process is documented in the Jupyter notebook “prepare_kinetic.ipynb” in the

folder “prepare_kinetic_data”. The solvent is parsed with SolventPicker object, where the influential

solvent is selected. The work to generate the solvent data is given in the Jupyter notebook

“Solvents.ipynb” in the folder “solvents”.

2.2 Input parsing and structure generation The input to the predict-SNAr workflow is the reaction SMILES, the reaction temperature in Kelvin and

the solvent name. The solvent is converted to SMILES via the Open Parser for Systematic IUPAC

nomenclature (OPSIN) tool (2.4.0).3,4 It is then checked against a list of solvents available in the

Gaussian16 program. If an exact match cannot be found, the most similar solvent based on distance in

the standardized three-dimensional space of dielectric constant (ε) and hydrogen bonding properties

(Abraham’s AH and BH) is used.5 If the hydrogen bonding parameters are not available for the input

solvent, the choice is based on just the dielectric constant. The same process is used for the xtb program,

which has a much more limited selection of solvents. Note that xtb energies are only used in

intermediate steps of the workflow and do not appear in the final output of the model. Therefore it does

not matter significantly what solvent is chosen for xtb step, as long as the right structures are found and

later optimized with Gaussian16.

The reaction SMILES is parsed first by the ReactionSmilesParser object which identifies substrate,

nucleophile, product and leaving group based on minimum common substructure matches (with some

additional rules pertaining specifically to the SNAr reaction). If the leaving group is not specified, it is

created. Intramolecular reactions are identified. Molecules which do not take part in the bond breaking

and bond making events are sorted as agents and placed “above the arrow” in the reaction SMILES (in

between the two “>>”). An AgentDetector object parses the agents to find acids and bases. The reactive

atoms are also identified.

The Smiles2XYZ object uses the output from the ReactionSmilesProcessor to construct 3D structures of

all reactants and products using the RDKit.6 The conformer generation recipe of Deane and co-workers

is used,7 with the ETKDG algorithm by Landrum and Riniker.8 Structures are optimized and ranked with

the MMFF9,10 or UFF11 force fields, depending on atom availability, keeping only the lowest-energy

conformer. Smiles2XYZ also constructs a reaction complex with the nucleophile situated at a distance of

6 Å from the substrate to serve as a starting point for further calculations with xtb. Smiles2XYZ further

identifies the atoms of the reactive ring and if the nucleophile and transition state would be candidates

for implicit/explicit solvation modelling (Section 2.3.5). The criterion for explicit solvation is that the

nucleophilic atom should be a negatively charged and of the second or third row of the periodic table.

The rationale is that these small anions should be more localized and harder for the implicit solvation

models to treat. Smiles2XYZ also detects the following cases which are of interest in the further

modelling:

1. Azide nucleophiles

2. Ortho nitro groups on the substrate

3. Non-conjugated nucleophilic atoms (adjacent to sp3-hybridized atom)

2.3 Mechanistic calculations

2.3.1 Quantum chemistry DFT calculations used Gaussian 16 Rev C.01.12 Geometry optimizations employed the ωB97X-D

functional13 with the 6-31+G(d)14,15 basis set as this functional has shown good performance for SNAr

reactions.16 All stationary points were confirmed with frequency calculations. Thermal contributions to

the free energies were corrected for low-frequency modes with Grimme’s quasi-harmonic scheme17

Page 25: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

4

using GoodVibes 3.0.118,19 at the reaction temperature. Final energies were obtained by combining

single-point energies with ωB97X-D/6-311+G(d,p)20 with the thermal free energies at the ωB97X-D/6-

31+G(d) level. Solvent effects were accounted for by the SMD solvation model.21 GFN2-xTB22 calculations

were performed with the xtb 6.2 software.23 Solvent effects with xtb were treated with the GBSA

solvation model. To better model anions, xtb calculations employed an electronic temperature (2000–

7000 K) to simulate the effect of diffuse basis functions (Section 2.3.6). Solvation of anionic nucleophiles

are generally treated poorly by continuum models, especially in polar solvents. To correct our solvation

energies, we used an automatized approximate version of the cluster-continuum model of Pliego and

Riveros (Section 2.3.5).24,25. For all structures, we performed conformational sampling with CREST (2.8)26

and xtb. The resulting conformers were ranked according to their electronic energy with ωB97X-D/6-

31+G(d) and only the lowest energy conformer was selected for further optimization.

2.3.2 Transition state calculations We first checked whether we could locate a stable σ complex to assess whether the reaction was

concerted or stepwise. If the σ complex was stable, we scanned each reactive bond separately using xtb,

starting from the σ complex, to find the TSs for addition and elimination. If the σ complex was not stable,

we performed a concerted scan of the reactive bonds using generalized internal coordinates (GICs) to

located the concerted TS, scanning from the reactive complex to the product complex. Single point DFT

calculations with ωB97X-D/6-31+G(d) along the scan coordinate identified the approximate location of

the TS, which was conformationally sampled with frozen reaction core and subsequently fully optimized

with DFT. TS conformations were sampled with atom position constraints on the aromatic core and bond

constraints for the reactive bonds. The reactions involving aliphatic alkoxides as leaving groups had

unrealistically large activation energies owing to proton-transfer being involved in the rate-determining

step. Our workflow treats this proton transfer as going from the nitrogen to the oxygen in a concerted

manner, while in reality it would occur by two separate proton transfers involving the solvent.27 Owing

to the inadequacy of our model, we use the barrier from the addition step as rate-determining in the

later machine learning, in accordance with the literature.28 Proper treatment of these types of

mechanisms will be the topic of a future study.

Reactant and product complexes were optimized with GFN2-xTB at a C–Nu distance corresponding to

the sum of the vdW radii of the two reactive atoms. The distance to the ortho carbons were also frozen

at distances determined from the C–Nu and C–ortho C distances together with the Pythagorean

theorem. For intermediate optimizations, the xtb optimization and CREST conformational search used

frozen C–LG and C–Nu bond lengths which were taken from the GFN2-xTB geometries of the substrate

and the product, respectively. For finding transition states, the GFN2-xTB energy profile was analysed

with respect to peaks higher than 0.01 kcal/mol, which were identified as candidate transition states. A

bond order criterion discarded peaks for which the bond order of any of the reactive bonds changed by

less than 0.05 Å. If no peak could be identified, the workflow moved on to the next stage, except in the

case of the second step of a stepwise mechanism, where the first geometry on the bond scan was taken

as a TS guess. The rationale was that the barrier could be very small and the peak had probably been

missed by the bond scan procedure. The peaks were sorted with respect to energy and an attempt to

locate the TS from the geometry of the first peak was attempted. If a TS could not be found, the program

proceeded to the next peak in the list. In the special case where concerted bond breaking and proton

transfer could occur in the second step of a stepwise mechanism, a GIC scan was conducted to find

additional guess transition states of this type if no other TS could be found. Transition states were

validated by projecting the normalized Cartesian coordinate displacements from Gaussian16 of the TS

mode onto the reactive bonds. Any TS candidate with a projected displacement below 0.13 units were

rejected. TS structures were optimized by DFT after the initial CREST conformational search as described

Page 26: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

5

above. First, the geometry was optimized with the reactive bonds frozen, and then a full TS optimization

was conducted.

For molecules including iodine, we used an effective core potential (ECP). The solution was to combine

the def2-SVPD and def2-TZVPD basis sets29,30 and their associated ECPs together with the 6-31+G(d) and

6-311+G(d,p) basis sets for the rest of the atoms, for double and triple zeta basis set calculations,

respectively. Generation of the basis set dictionaries is documented in the Jupyter notebook

“create_pickle_files.ipynb” in the directory “basis sets”.

2.3.3 Use of additional constraints for xtb optimizations We used some additional constraints for xtb calculations to avoid complications were xtb would favour

geometries that would be very high in energy at the DFT level or lead to discontinuities in bond scans.

These constraints were lifted for the final DFT optimization and do not impact the final geometries that

are used to calculate the reaction barriers and the descriptors.

We found that the combination of ortho nitro groups on the substrate together with amine nucleophiles

could cause problems with xtb optimizations using an electronic temperature. During optimizations and

conformational sampling of intermediates and transition states, spurious proton transfer could happen

from the amine to the nitro group. Therefore, we froze the N–H bonds for this type of calculations to

prevent the spurious transfer.

Azide nucleophiles tended to bend at the GFN2-xTB level in intermediates and transition states, which

could cause problems with discontinuities in the DFT single point calculations along the GFN2-xTB bond

scans. We therefore imposed two constraints for these cases (Figure S2):

1. The N–N–N angle is frozen to 180 degrees (between atoms 15, 16 and 17 in the figure)

2. The Lg–C–N–N dihedral angle is frozen at 180 degrees (between atoms 10, 3, 17 and 16 in the

figure).

Figure S2. Constraints for azides used in xtb optimizations and conformational samplings. Angle constraint of 180 degrees for the angle 15-16-17 and a dihedral angle constraint of 180 degrees for 10-3-17-16.

For bond scans, we also implemented an additional set of constraints. For scans, for the first step of a

step-wise mechanism, we implemented constraints for negatively charged oxygen nucleophiles which

were connected to saturated atoms. Due to lack of diffuse functions, the O–C bond is artificially short

with GFN2-xTB due to negative hyperconjugation effects to relieve the instability of the negative charge

on the oxygen atom. We therefore froze the C–O distance during the scan to the value calculated with

DFT for the isolated nucleophile.

Page 27: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

6

For all the GIC bond scans, we also froze the difference of the distances between the nucleophilic atom

and the ortho carbons of the substrate to ensure a symmetric approach of the nucleophile. For the bond

scans from the intermediate with xtb, we instead scanned the Nu–ortho C distances together with the

C–Nu distance.

2.3.4 Calculation monitor The calculation monitor handles common errors and issues that occur during electronic structure

calculations and geometry optimizations.

Displacement of negative frequencies for geometry optimizations

If a minimum structure is optimized and the frequency calculation shows one or more negative

frequency, the structure was displaced along the corresponding normal mode and the optimization

restarted. For the workflow, one such preoptimization was done.

Dissociation check

The calculation monitor checked for dissociation of bonds in the optimization of the intermediate.

Dissociation was judged to have occurred if the bond orders of either the C–Nu or C–Lg bond decrease

below 0.3. The calculation was then aborted and the workflow proceeded to calculate a concerted

mechanism.

Adaptive keyword changes in response to SCF errors and coordinate system errors

Common convergence errors are captured and appropriate Gaussian keywords are inserted to try to

remedy the situation. One example is the error "Convergence failure -- run terminated" for which the

keyword “scf=xqc” is used. Coordinate system errors are also addressed, e.g., “RedCar failed” which is

addressed with a series of Gaussian IOps: "1/59=10", "1/59=14", "1/59=4", "1/59=40" and "1/59=44".

Also more obscure and non-reproducible errors such as “Internal input file was deleted!” are handled by

simply restarting the calculation.

2.3.5 Explicit solvation Explicit solvation is handled via the cluster-continuum model of Rivero and Pliegos.25 We targeted the

energy that is missed by the implicit solvation model, which can be corrected by using explicit solvent

molecules. We denote the free energy in solvent according the full cluster-continuum model with n

explicit solvent molecules as ΔGsolv,n(solute), while that using the implicit model is ΔGsolv(solute). The

correction to the implicit model is then given by

Δ𝐺𝑠𝑜𝑙𝑣,𝑛(𝑠𝑜𝑙𝑢𝑡𝑒) − Δ𝐺𝑠𝑜𝑙𝑣(𝑠𝑜𝑙𝑢𝑡𝑒) = Δ𝐺𝑠𝑜𝑙𝑣(𝑐𝑙𝑢𝑠𝑡𝑒𝑟) − Δ𝐺𝑠𝑜𝑙𝑣(𝑠𝑜𝑙𝑢𝑡𝑒) − 𝑛Δ𝐺𝑠𝑜𝑙𝑣(𝑠𝑜𝑙𝑣𝑒𝑛𝑡) Equation 1

where ΔGsolv(cluster) is the free energy of the cluster, ΔGsolv(solute) the free energy of the solute and

ΔGsolv(solvent) is the free energy of the solvent (with the appropriate standard state), using the implicit

solvation model. The number of solvent molecules, n, is determined by a variational principle, where the

n which gives the largest correction is the most appropriate. This variational principle stems from two

competing factors. The first factor is the stabilizing effect due to specific interactions between the

explicit solvent molecules and the solute, which are not captured fully by the implicit solvation model.

The second factor is the destabilizing effect of bringing solvent molecules from the bulk solvent to form

the cluster, which corresponds to an entropic loss.

We have implemented an approximate version of the cluster-continuum model that uses a combination

of GFN2-xTB and DFT to select the optimal number of solvent molecules (Figure S3). The final correction

term is then calculated with full DFT.

Page 28: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

7

Figure S3. Sketch of automated cluster generation process.

The first step of the procedure is to generate a cluster geometry. We start by placing solvent molecules

around the solute at a distance of 4 Å at the geometries given in Table S1.

Table S1. Geometry of initial placement of solvent molecules.

n Geometry

1 Point 2 Line 3 Equilateral triangle 4 Tetrahedron 5 Trigonal bipyramid 6 Octahedron

The cluster is the optimized with a small force constant (0.0005) pulling the solvent molecule towards

the closest atom in the molecule, to form a crude guess for the cluster geometry. The pulling step is

followed by two rounds of CREST conformational searches in the non-covalent interactions (NCI) mode.

The conformers generated in the second run are ranked with DFT, and cluster free energy is constructed

from the DFT single-point energy together with the free-energy contributions from GFN2-xTB. The

energy of the solvent and solute are calculated in the corresponding manner, and the solvent correction

term is calculated from Equation 1. Solvent molecules are added iteratively until the solvent correction

terms decreases in magnitude according to the variational principle. A cut off of 2.0 kcal/mol is used to

decide whether to go ahead with the final DFT cluster optimization. Special treatment is done for

hydroxide in water, where GFN2-xTB optimization gives a delocalized structure where the proton is

shared equally between two O atoms (

Figure S4). In this case, the O–H bond lengths are frozen for all steps. For transition states, the aromatic

core and the reactive atoms are frozen. For the final DFT optimization of the cluster at the optimized

number of solvent molecules, all constraints are released. For transition states, this means a full TS

optimization for the cluster. The final solvent correction at the DFT level is then again calculated

according to Equation 1.

Page 29: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

8

Figure S4. Hydroxide with one explicit water molecule optimized with GFN2-xTB

2.3.6 Use of electronic temperature in xtb calculations The xtb program uses a standard electronic temperature of 300 K. We found that a higher electronic

temperature could serve as a substitute for diffuse functions for the GFN2-xTB method, which lacks them

by default. The rationale is that electrons of the approaching anionic nucleophile could partly delocalize

into the π* orbitals of the aromatic ring. However, if the delocalization becomes too strong, complete

charge transfer is the result. It is therefore necessary to regulate the electronic temperature carefully to

keep in the window where the lack of diffuse functions is remedied, while the problems of charge

transfer don’t become too apparent. We found that a good starting guess could be based on the energy

gap between the HOMO energy of the nucleophile and the LUMO energy of the substrate. The following

rules were used to set the initial electronic temperature, which could later be changed for some steps.

HOMO-LUMO gap (eV) Electronic temperature (K)

(−∞) – (−0.5) 4000 (−0.5) – (+∞) 7000

After the initial setting based on the separated reactants, the following rules were used, which reflect

that the HOMO-LUMO gap of the reaction complex is different than for the separated reactants.

HOMO-LUMO gap (eV) Electronic temperature (K)

(−∞) – (+2.0) 4000 (+2.0) – (+∞) 7000

Bond scans and conformational scans are examples were the workflow will recheck the HOMO-LUMO

gap and adjust the electronic temperature. For bond scans with neutral nucleophiles, the following

ranges are used.

HOMO-LUMO gap (eV) Electronic temperature (K)

(−∞) – (+1.0) 2000 (+1.0) – (+2.0) 4000 (+2.0) – (+∞) 7000

2.4 Feature generation The solvent-accessible surface area and the Pint dispersion descriptor were calculated with an in-house

development version of the morfeus software.31 The SASA calculation used the algorithm by Shrake and

Rupley32 with the vdW atomic radii taken from the CRC Handbook.33 Pint was calculated on a surface

constructed from the atomic radii of Rahm and co-workers34 and with the D3 dispersion coefficients.35

Electronic structure features were calculated based on B3LYP/6-31+G(d) calculations with the SMD

solvation model. The local electron attachment energy36 and the average local ionization energy37 as

well as the surface electrostatic potential were calculated with the HS95 program (version 190510).38

The Is,min and Es,min values were taken as the lowest values on the atomic surface, regardless if there was

a stationary point associated with the atom in HS95. The five solvent PCA components compiled by

Diorazio et al. were used as solvent features.5 DDEC6 charges and bond orders were calculated with

Page 30: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

9

Chargemol (3.5)39,40 The electrostatic potential at the nuclei (VN) was calculated with Gaussian 16 using

the keyword prop=potential. The global nucleophilicity parameter N was calculated as the negative of

the ionization potential of the substrate in solution.41 The global electrophilicity parameter ω was

calculated as

𝜔 =𝜇2

2𝜂 Equation S2

where η is the chemical hardness given by

𝜂 = 𝐼𝑃 − 𝐸𝐴 Equation S3

Here, IP and EA are the vertical ionization potential and electron affinity, respectively, and μ is the

chemical potential,42 given by

𝜇 = −𝐼𝑃 + 𝐸𝐴

2 Equation S4

The local electrophilicity descriptor was calculated as

𝑙𝜔 = −𝜇

 𝜂𝑓 +

1

2(𝜇

𝜂)2

𝑓2 Equation S5

where f is the quadratic Fukui function, and f2 is the dual descriptor.43 The local nucleophilicity descriptor

was calculated as

𝑙𝑁 = 𝑓− Equation S6

where f− is the Fukui function for electrophilic attack. The quadratic Fukui function was calculated as

𝑞−+𝑞+ 

2 Equation S7

where q− is the charge of the atom in the anion, and q+ is the charge of the atom in the cation. The dual

descriptor was calculated as

𝑓2 = 𝑓+ + 𝑓− Equation S8

where f+ is the Fukui function for nucleophilic attack. The Fukui functions for nucleophilic and

electrophilic attack were in turn calculated as

𝑓+ = 𝑞 − 𝑞+ Equation S9

and

𝑓− = 𝑞− − 𝑞 Equation S10

where q is the charge of the neutral molecule. We used atomic Hirshfeld charges44 that were calculated

with Gaussian 16 using the population=hirshfeld keyword.

Morgan fingerprints were calculated with the RDKit (2020.03.1.0)6 as count vectors and length 1024 bits.

Reaction difference fingerprints were obtained by subtracting the summed fingerprints of the reactants

from the summed fingerprints of the products. Using bits instead of counts, or using a length of 512 or

2048 instead 1024 did not change the results significantly (Table S5).

Reactions were atom-mapped using Biovia Pipeline Pilot 2018 (18.1.0.1604),45 and erroneous mappings

were corrected by hand using ChemFinder (19.0.0.22) and ChemDraw (19.0.0.22). CGRs were generated

Page 31: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

10

with CGRtools (4.0.18)46,47 and were aromatized and standardized. ISIDA descriptors were then obtained

with CIMtools (4.0.2)48 and Fragmentor 2017. Sequence features were generated with the following

keywords: “fragment_type=3”, “min_length=2”, “max_length=8”, “cgr_dynbonds=1”,

“doallways=True”, “useformalcharge=True”. Atom features were generated with the following

keywords: “fragment_type=6”, “min_length=2”, “max_length=8”, “cgr_dynbonds=1”,

“useformalcharge=True”. The generation of the descriptors is documented in the Jupyter notebook

“cgr_tools.ipynb” in the folder “cgr_isida_descriptors”. One-hot encoded (OHE) features were created

using the full data set prior to the train-test split. The encoding was based on the identity of the

substrate, nucleophile, product, leaving group and solvent, as given by their InChIKeys. The BERT

reaction fingerprints49 (of fixed length 256 bits) were generated with the rxnfp (0.0.1) package.50 For

prediction, we used the pre-trained model (“bert_pretrained”), and for making reaction maps we used

the fine-tuned model (“bert_ft”). The generation of the BERT fingerprints is documented in the Jupyter

notebook “rxnfp.ipynb” in the folder “bert_rxnfp”.

2.5 Running the workflow The workflow is designed as a Python package predict_snar, version 0.1.0, that can be installed with

“pip install .”. The Python package requirements are specified in the “setup.py” file and listed in Table

S2.

Table S2. Python packages required to run the predict-SNAr workflow.

Requirement Used version Description

ase 3.18.0 Atomic simulation environment.51 Handling of structure information. cclib 1.6.2 Parsing of quantum-chemical chemistry output.52 Documentation.

joblib 0.13.2 Parallelization of calculations. Documentation. mendeleev 0.4.5 Periodic table information such as covalent radii and vdW radii. Documentation. numpy 1.16.4 Array manipulation and mathematical calculations. Documentation. goodvibes 3.0.1 Corrections to quantum-chemical free energies. Documentation. scipy 1.3.2 Constants and unit conversion. Distance calculations. Peak finding. steriplus 0.3.0 Calculation of Pint and SASAr descriptors. Available in the near future as morfeus.

The workflow is designed for running on a Linux cluster. Required software is listed in Table S3.

Table S3. Software required to run the predict-SNAr workflow.

Software Description

Gaussian16 Quantum-chemistry. Commercial license. HS95 Electronic descriptors. Needs permission from Tore Brinck Chargemol Charge and bond order descriptors. Install from here. interface_script A Python script to handle communication between xtb and Gaussian (included) xtb Semi-empirical quantum chemistry. Install from here. CREST Conformational sampling. Install from here.

The workflow has been tested with the versions of software noted in Section 2.3. A configuration file

called “.ps_config” must be placed in the user home folder with desired default options and directories

to software. Java must also be available in the environment to run the packaged OPSIN tool.

[DFT]

solvation_model = smd

sp_solvation_model = smd

nosymm = True

[DIRECTORIES]

xtb = /projects/cp/programs/xtb/default # Path to xtb executable

Page 32: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

11

crest = /projects/cp/programs/crest/default # Path to crest executable

chargemol = /projects/cp/programs/chargemol/default # Path to chargemol executable

hs95 = /projects/cp/programs/hs95/default # Path to hs95 executable

interface_script = /projects/cp/programs/scripts/predict-snar/interface-g16-xtb # Path to interface script executable

crest_scratch = /dev/shm/ksrf385 # Scratch directory for CREST program

gaussian_scratch = /scratch/ksrf385 # Scratch directory for Gaussian program

atomic_densities = /projects/cp/programs/chargemol/default/atomic_densities/ # Directory to atomic densities used by the chargemol program.

A config file must be created in the directory before running the workflow. This is done by running the

script ps_create_config. For instructions on its use, run ps_create_config --help.

The workflow is invoked with the command

python -m predict_snar -p <n_cpus> -m <mem in GB> ‘<reaction SMILES>'

where the argument “-p” specifies the number of CPUs to use and “-m” the amount of available memory

in GB. An example job script is included in the supporting information with the article. Two example

outputs from the workflow are also given.

3 Machine learning The machine learning is reproduced in the Jupyter notebooks listed in Table S4.

Table S4. Jupyter notebooks used for machine learning.

Notebook file Description

“train_test_split.ipynb” Data pre-processing, feature generation, train-test split. “modelling.ipynb” Model selection “testing.ipynb” Validation on external test set. Dataset visualization. Feature importances. Learning curves.

3.1 Model selection For the model building, we used the Python machine learning package scikit-learn (0.22.1).53 Data was

handled by the pandas (1.0.3) data analysis and manipulation package.54 Plots were generated with

matplotlib (3.1.0)55 and seaborn (0.9.0).56 We first processed the dataset in accordance with the

description in Section 5.1 and put the reactions through the workflow(Section 2). We then split the

dataset into a training set (80%) and a test set (20%) and proceeded with model selection on the training

set. As the first step in the model selection, we choose between different variations of the DFT activation

energies, seeing that the best performer was with full cluster-continuum solvent treatment of both

nucleophile and TS, using the Grimme quasi-harmonic correction for the free energies (Figure S5). We

investigated correlation between the features using the Pearson57 correlation matrix and the variance

inflation factors.58 We confirmed the initial feasibility of modelling using linear regression and inspecting

the normal probability plot (Q-Q plot) of the residuals, with reassuring result.

Page 33: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

12

(a) Cluster-continuum solvation of nucleophile and TS

(b) Cluster-continuum solvation of nucleophile

(c) No cluster-continuum solvation

Figure S5. Correlation between DFT-computed activation free energy and experimental activation free energies with cluster-continuum solvation of (a) nucleophile and TS, (b) only nucleophile and (c) no molecules. y: ΔG‡

Exp (kcal/mol), ŷ: ΔG‡DFT (kcal/mol)

We used the bias-corrected bootstrap cross-validation (BBC-CV) method for model selection, with the

goal to avoid doing nested cross-validation. BBC-CV is a more economical alternative to nested cross-

validation that allows unbiased comparison between different families of algorithms.59 It can also

estimate the selection bias for choosing the best estimator based on the cross-validation scores.

Empirical studies have found this estimate to give slightly worse scores as compared to the final

evaluation on an external test set.60 For methods where hyperparameter tuning was done via grid search,

all resulting models were grouped within a family (e.g., random forest, SVR etc.), and BBC-CV was used

to estimate the bias due to overfitting on the cross-validation procedure within this family.61 This bias

was then subtracted from the score of the best model in the family before the scores of each family were

Page 34: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

13

compared. Hyperparameter tuning was done using grid search, with the GridSearchCV object in scikit-

learn, picking the model with the best performance within each family of models. All models were

evaluated on the same 10 cross-validation folds, which were generated at the beginning of the BBC-CV

procedure. To create interaction features and polynomial features for linear models, we used the

PolynomialFeatures object in scikit-learn with order to second order. Standardization was applied to

numeric features, and excluded categorical or count features such as fingerprints.

The full results showed that the GPRM3/2 was the strongest contender on the MAE score and we therefore

chose it as our final model to be evaluated on the external test set (Table S5).

Table S5. Results for machine learning models. Uncertainty intervals are given as one standard error of the mean. Abbreviations explained in Section 3.1.1. Best method for each score marked in bold.

Method Feature set R2 (kcal/mol) MAE (kcal/mol) RMSE (kcal/mol)

Baseline

DFT – 0.09 2.93 3.73 DFT linear fit – 0.63−0.03

+0.03 1.74−0.08+0.07 2.36−0.16

+0.15 Mean – – 2.84−0.13

+0.13 3.90−0.25+0.24

Median – – 2.81−0.13+0.13 3.93−0.24

+0.24

Machine learning models

Linear regression Full 0.79−0.03+0.03 1.20−0.06

+0.06 1.77−0.14+0.13

KNN Full 0.74−0.03+0.04 1.37−0.07

+0.07 1.95−0.12+0.08

Bayesian ridge Full 0.77−0.03+0.03 1.24−0.07

+0.06 1.83−0.14+0.14

PF2 0.69−0.06+0.09 1.23−0.10

+0.08 2.16−0.26+0.24

PLS Full 0.79−0.02+0.03 1.25−0.06

+0.06 1.79−0.13+0.13

PF2 0.84−0.02+0.03 0.98−0.05

+0.05 1.54−0.12+0.12

ARD Full 0.79−0.02+0.03 1.21−0.06

+0.06 1.76−0.11+0.11

PF2 0.85−0.02+0.03 0.89−0.05

+0.05 1.46−0.13+0.14

Random forest Full 0.84−0.02+0.02 0.98−0.06

+0.05 1.53−0.10+0.11

Gradient boosting Full 0.84−0.02+0.03 1.00−0.06

+0.06 1.57−0.13+0.12

SVR Full 0.85−0.02+0.02 0.81−0.06

+0.06 1.48−0.15+0.15

GPR M 3/2 Full 𝟎. 𝟖𝟕−𝟎.𝟎𝟐±𝟎.𝟎𝟐 𝟎. 𝟖𝟎−𝟎.𝟎𝟔

+𝟎.𝟎𝟔 1.41−0.14+0.14

GPR M 5/2 Full 𝟎. 𝟖𝟕−𝟎.𝟎𝟐+𝟎.𝟎𝟑 0.81−0.06

+0.05 1.40−0.13+0.14

GPR RQ Full 𝟎. 𝟖𝟕−𝟎.𝟎𝟐+𝟎.𝟎𝟑 0.83−0.06

+0.05 1.41−0.13+0.13

GPR RBF Full 𝟎. 𝟖𝟕−𝟎.𝟎𝟐+𝟎.𝟎𝟐 0.83−0.05

+0.05 1.40−0.14+0.14

Reduced feature sets

GPR M 3/2 Small 0.84−0.02+0.03 0.88−0.06

+0.06 1.53−0.14+0.15

No TS 0.86−0.02+0.02 0.86−0.05

+0.05 1.44−0.12+0.14

Only TS 0.83−0.03+0.03 1.03−0.06

+0.06 1.68−0.13+0.14

Surface 0.80−0.03+0.03 1.10−0.07

+0.07 1.76−0.14+0.15

Traditional 0.81−0.03+0.04 1.00−0.06

+0.06 1.68−0.15+0.16

Structural information

GPR M 3/2 OHE 0.62−0.04+0.04 1.58−0.09

+0.09 2.39−0.23+0.23

ISIDA atom 0.53−0.06+0.07 1.64−0.10

+0.10 2.66−0.26+0.27

Morgan1 0.68−0.03+0.03 1.55−0.08

+0.08 2.20−0.13+0.12

Morgan2 0.74−0.03+0.03 1.34−0.07

+0.07 1.94−0.15+0.17

Morgan3 0.80−0.03+0.03 1.09−0.06

+0.06 1.72−0.19+0.24

Morgan4 0.79−0.03+0.03 1.13−0.07

+0.07 1.76−0.18+0.22

Morgan5 0.79−0.03+0.03 1.15−0.07

+0.07 1.81−0,19+0.23

Morgan6 0.78−0.03+0.03 1.19−0.07

+0.07 1.81−0.18+0.23

ISIDA seq 0.69−0.03+0.04 1.37−0.08

+0.08 2.15−0.14+0.13

BERT pre-trained

0.85−0.02+0.02 1.03−0.05

+0.05 1.51−0.10+0.10

BERT fine-tuned 0.77−0.03+0.03 1.29−0.06

+0.06 1.82−0.09+0.10

Structural + physical organic

GPR M 3/2 Full + Morgan 3 0.86−0.02+0.03 0.87−0.06

+0.05 1.43−0.13+0.12

Page 35: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

14

GPR M 3/2 Full + BERT pre-trained

𝟎. 𝟖𝟕−𝟎.𝟎𝟐+𝟎.𝟎𝟑 0.86−0.05

+0.05 𝟏. 𝟑𝟒−𝟎.𝟏𝟐+𝟎.𝟏𝟏

3.1.1 Models tested Support vector regression (SVR).62 We used the RBF kernel and features were standardized. A grid search

was carried out to optimize the values of the kernel coefficient γ and the regularization parameter C,

both with the same range of values (0.001000, 0.004642, 0.02154, 0.1000, 0.4642, 2.154, 10.00, 46.42,

215.4, 1000).

Random forest regression (RF).63 The number of trees (10, 20, 100, 200) and maximum tree depth (5, 10,

15, 20, None) were optimized.

Gradient boosting regression (GB).64 We used 1000 boosting stages and early stopping with a threshold

of 10 iterations without improvement. A grid search optimized the learning rate (0.001, 0.01, 0.1) and

maximum depth of each individual tree estimator (1, 2, 3, 4, 5).

Gaussian Process Regression (GPR).65 We scaled both the features and the target using the

TransformedTargetRegressor in scikit-learn. We tried the radial basis function (RBF), rational quadratic

(RQ) and Matern 3/2 and 5/2 (M 3/2, M 5/2) kernels. The final composite kernel was created by

multiplying with a constant kernel and adding a white kernel to model the noise.

Bayesian ridge regression (BRR).66 This Bayesian version of ridge regression tunes the regularization

parameter automatically from the data.

Automatic Relevance Determination (ARD).67 ARD produces a more sparse solution than BRR as is

suitable when a large number of features are used.

Partial Least Squares (PLS).68 A linear method that can handle multicollinearity and a large number of

features compared to samples. We used a grid search between 1 and 15 to select the number of

components.

K-nearest neighbours regression (KNN).69 Serves as a non-parametric baseline method.

Dummy regression models. We used the mean and the median of the training set for future prediction

as reasonable baseline models. These models were implemented with DummyRegressor in scikit-learn.

3.2 Dataset visualization The full feature set Xfull was decomposed with Principal components analysis (PCA)70 and the two

components with highest explained variance were used for the visualization. For the PLS + UMAP plots,

we first used PLS for supervised dimensionality reduction. The number of components were chosen

based on the R2 score, using the one-standard error rule to select the model with fewest dimension with

a score within one standard error of the best-scoring model overall. This approach led to a model with 5

components. The Xfull feature space was transformed into this five-dimensional latent space of the PLS

model, and further reduced to two dimension with the Uniform Manifold Approximation and Projection

(UMAP) method,71 using the UMAP python package,72 with standard settings, except the keyword

min_dist=0.3, and n_neighbors=10. A reaction map (Figure S6) was generate with the tmap package

(1.0.4)73,74 and visualized with faerun (0.3.10).75,76 The Jupyter notebook “reaction_map.ipynb” in the

folder “reaction_map” gives the details of how this map was generated. An interactive version of the

reaction map is given in the file “index.html”.

Page 36: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

15

Figure S6. Reaction map generated with XBERT, tmap and faerun. Annotations of nucleophile and leaving group types has been done manually.

3.3 Feature importances Before assessing the feature importances, we checked the variance inflation factors (Figure S7–Figure

S9) and Pearson correlation matrices (Figure S10–Figure S12) to assess the extent of collinearity among

the features in Xfull and XnoTS and Xsmall. It is clear that Xfull and XnoTS have large problems with collinearity,

especially in comparison with Xsmall. We therefore decided to cluster the features with SciPy (1.2.1)77

based on the Spearman rank correlation,78 using the Ward linkage and the distance criterion (following

a recipe from the scikit-learn documentation). The threshold for clustering was selected manually to

reduce the number of highly correlated features while retaining a clear chemical interpretation of each

cluster (Figure S13–Figure S15). We computed the permutation importances version of feature

importances with the GPRM3/2 model together with the permutation_importances function from scikit-

learn. We used 10 repeats, and the MAE score and the stability of the feature ranking was assessed by

running 10 bootstrap samples with BootstrapOutOfBag in mlxtend.79

Page 37: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

16

Figure S7. VIFs for Xfull. A value above 10 is indicative of high collinearity with other features. Note log scale.

Figure S8. VIFs for XnoTS. A value above 10 is indicative of high collinearity with other features. Note log scale.

Page 38: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

17

Figure S9. VIFs for Xsmall. A value above 10 is indicative of high collinearity with other features. Note log scale.

Figure S10. Pearson correlation matrix for Xfull.

Page 39: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

18

Figure S11. Pearson correlation matrix for XnoTS.

Figure S12. Pearson correlation matrix for Xsmall.

Page 40: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

19

Figure S13. Feature clustering dendrogram for Xfull.

Figure S14. Feature clustering dendrogram for XnoTS.

Page 41: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

20

Figure S15. Feature clustering dendrogram for Xsmall.

3.4 Learning curves Learning curves were calculated with the learning_curve function in scikit-learn with splits of 10%, 20%,

30%, 40%, 50%, 60%, 70%, 80%, 90% and 100% of the data using 10-fold cross-validation. To assess how

much data is needed to accurately predict the DFT activation energy, we used XnoTS as the feature set.

3.5 Analysis of reactions with large errors Reactions with absolute residuals larger than 2.0 kcal/mol for the training and test sets with the GPRM3/2

model are given in Table S6 and Table S7, respectively.

Table S6. Reactions with error larger than 2.0 kcal/mol in the training set.

Reaction Residual error (kcal/mol)

−5.00

−2.90

5.10

4.23

−2.25

Page 42: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

21

−3.50

5.52

2.27

4.41

−2.41

−2.58

Table S7. Reactions with error larger than 2.0 kcal/mol in the test set.

Reaction Residual error (kcal/mol)

−2.11

−2.22

−2.50

2.06

−2.48

Page 43: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

22

−2.39

−2.52

2.27

3.6 Dependence on number of heavy atoms We checked the dependence of the model error on the number of heavy atoms to see if more complex

molecules would display larger error, but could see no such trend (Figure S16).

(a) (b)

Figure S16. (a) Distribution of number of heavy atoms in the data set. (b) Mean absolute error for test set depending on the number of heavy atoms.

3.7 Y-randomization We further tested for overfitting by y-randomization, in which the link between the feature vectors in X

and the experimental activation free energies in the response vector y is broken by randomizing the y

vector (Table S8).80 The R2 for GPRM1.5 is lowered from 0.86 to −0.05, indicating complete lack of learning

when y is randomized. This is strong evidence that the models are not just learning noise but are finding

a real signal in the feature data. We performed the y-randomization using the permutation_test_score

function in scikit-learn with 10 permutations and 10-fold cross-validation for each permutation, as

recommended in the literature.81

Table S8. Results of y-randomization with GPRM1.5 on Xfull for the training set.

R2 MAE (kcal/mol) RMSE (kcal/mol)

True score 0.86 0.79 1.35 y-randomized score −0.05 2.87 3.88

4 Regio- and chemoselectivity validation We used a reaction dataset from the patent literature, collected by Landrum and co-workers.82 The

dataset (“dataSetB.csv”) comprised 50,000 reactions that were first classified with the NameRxn (3.1.9)83

tool and then filtered according to reaction classes that include SNAr reactions according the SNAr

Page 44: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

23

“concept” provided by NextMove (Table S9). Reactions with transition metals were excluded as they

would likely go via the Buchwald-Hartwig reaction. The dataset was further filtered manually to remove

reactions which apparently did not go via the SNAr mechanism (e.g., SN2 reactions). Reactions including

boronic acids or phosphine ligands (presumed to occur via Buchwald-Hartwig) were also removed.

Furthermore, some reactions also included the product among the starting material and were removed.

Table S9. Reaction types including in the SNAr concept used to filter out SNAr reactions from the patent data.

Code Reaction name

1.3.1 Bromo Buchwald-Hartwig amination 1.3.2 Chloro Buchwald-Hartwig amination 1.3.3 Iodo Buchwald-Hartwig amination 1.3.4 Triflyloxy Buchwald-Hartwig amination 1.3.5 Chan-Lam arylamine coupling 1.3.6 Bromo N-arylation 1.3.7 Chloro N-arylation 1.3.8 Fluoro N-arylation 1.3.9 Iodo N-arylation 1.3.10 Triflyloxy N-arylation 1.3.11 Chichibabin amination 1.3.12 Mesyl N-arylation 1.3.13 Mesyloxy N-arylation 1.3.14 Tosyloxy N-arylation 1.7.11 SNAr ether synthesis 9.7.96 Sandmeyer bromination 9.7.97 Sandmeyer chlorination 9.7.98 Sandmeyer fluorination 9.7.99 Sandmeyer iodination 1.1.2 Menshutkin reaction 1.8.5 Thioether synthesis 9.7.106 Fluoro to azido 9.7.166 Formyl to cyano 9.7.39 Chloro to amino 9.7.44 Chloro to hydroxy 9.7.64 Fluoro to amino

The pruned set of reactions comprised 4353 SNAr reactions, of which 1209 had alternative reactive sites

with halogens on the reactive ring, leading to questions relating to regio- and chemoselectivity (Figure

S17). These reactions were sorted according to the molecular weight of the product, and all possible

reactions relating to regio- and chemoselectivity of the reactive ring were generated for the 100

reactions with lowest product molecular weight. Sorting on molecular weight was done to allow a quickly

calculated validation test.

Figure S17. Two examples of reactions with regio- and chemoselectivity questions.

The protonation state of the nucleophile was not included in the patent data, and needed to be set. We

used the following set of rules:

1. –OH or –SH nucleophile: Deprotonate

2. Aromatic –NH nucleophile part of diazole or triazole ring:

a. If strong base present (pKa > 10): Deprotonate

b. Else: The reactive tautomer is where the nucleophilic atom doesn’t have an H

Page 45: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

24

3. Aromatic –NH nucleophile not part of diazole or triazole ring: Deprotonate

4. Non-aromatic -NH:

a. If strong base present: Deprotonate

b. Else: Neutral –NH acts as nucleophile.

For a negatively charged nucleophile, the leaving group was also considered to be negatively charged.

As reaction solvent and temperature was not available in the dataset, we used a set of standard reaction

conditions: acetonitrile for neutral reactions and methanol for ionic reactions and a reaction

temperature of 298 K. This is expected to introduce some noise in the predictions. Solvents are

sometimes included in the reaction SMILES in the dataset, but it is not always clear what is the reaction

solvent and what was used for, e.g., work-up.

For analysis of the results, we constructed the reaction feature matrix Xfull for the validation reactions

and used the previously trained GPRM3/2 model for prediction. The isomer corresponding to the reaction

with the lowest activation energy was considered as the predicted major product. For comparison, we

also used the DFT activation energies.

The validation work is documented in the notebooks listed in Table S10. The validation database is given

in the folder “validation_2020_07_04”.

Table S10. Jupyter notebooks used for validation work under the folder “validation_dataset”.

Notebook file Description

“validation_data.ipynb” Generation of the 100 reactions. “prediction.ipynb” Prediction with ML method and DFT vs. experiment.

5 Experimental dataset

5.1 Description of the modelling data An extensive literature search was performed to identify intermolecular nucleophilic aromatic

substitution reactions with associated second-order kinetic data with a focus on reactions where the

initial nucleophilic attack is rate-limiting. The reaction dataset was extracted from 37 references

spanning the years 1950–2019. Kinetic data is not curated by the standard reaction databases, and thus

manual extraction and curation was required. In total 518 data points were extracted (the full dataset),

and 446 where used in the model building (the modelling dataset). The reason for this smaller number

is that 48 of the data points were found after the modelling was finished, 15 had missing data which

prevented them from being used, 6 failed in the transition state calculations, two were carried out in a

solvent (tetramethylsilane) for which we did not have solvent features, and one was removed due to a

mismatch in the reaction SMILES during the modelling process. The whole data set (518 points) covers

88 different nucleophiles, 121 different substrates and 41 different solvents. The experimental

temperature range is 273–468 K and log k, the base-10 logarithm of the second-order rate constant,

spans −15.9 to −3.59. Rate constants were converted to activation energies at the temperature in

question using the Eyring equation

Δ𝐺‡ = −𝑅𝑇 ln𝑘ℎ

𝑘𝑏𝑇 Equation S11

where R is the gas constant, T is the reaction temperature, k is the second-order rate constant, h is the

Planck constant, kb is the Boltzmann constant and using a standard state of 1 M. Of the reactions

reported, 284 have one set of reaction conditions and 70 reactions have more than one set of reaction

conditions reported. The partial dataset (446 reactions) used for modelling is described in the main

article. A record of this database is given in the files “SNAR_reaction_dataset_SI.csv” and

Page 46: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

25

“SNAR_reaction_dataset_SI.xlsx”. Each dataset record includes the following information: canonicalized

SMILES of the reaction, reactants and products, second-order rate constants, activation free energy,

temperature, solvent, literature source information and flag on its use in the model building process. A

full listing of the column names and a corresponding description is given in Table S11.

Table S11. Definitions of columns in full reaction dataset.

Column Description

Entry Unique identification number of reaction Exp_Rate_Constant k1 (M-1s-1) Experimental second-order rate constant reported in M-1s-1 Substrate SMILES Canonicalized SMILES of substrate Nucleophile SMILES Canonicalized SMILES of nucleophile Product SMILES Canonicalized SMILES of product Reaction SMILES Canonicalized SMILES of reaction Temp (K) Reaction temperature in Kelvin Activation Free Energy (kcalmol-1) Activation Free energy in kcalmol-1 calculated using the

experimental rate constant and the Eyring equation Reference Journal reference for experimental data Year Year of reference Title Title of reference Authors Authors of reference DOI Reference digital object identifier Book_Page Page from referenced book where experimental data is extracted

from Table Table in reference the experimental data is extracted from Nucleophile Name of nucleophile Leaving group Name of leaving group Solvent Solvent for reaction reported. logk Base-10 logarithm of second-order rate-constant. Entry IDs of duplicates Entry identifiers for duplicate reactions without considerations for

reaction conditions. Only four examples with identical reaction conditions: 223 and 468, 27 and 109, 140 and 490, 220 and 469.

Model Flag “modelled”: Used in the machine learning model. 443 unique reactions with 3 additional replicates. “not modelled”: Reactions that were extracted and added to the data set after model building was completed. “failed”: Reactions that failed the predict-SNAr workflow due to TS finding problems (not finding or finding wrong TS) leading to no barrier, very low barrier, or very high barrier. “removed”: Reactions that were removed for other reasons. “missing data”: Reactions that lack activation energies, temperature or solvent.

5.2 Distribution of activation free energies The distribution of activation free energies in the modelling dataset is given in Figure S18.

Page 47: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

26

Figure S18. Distribution of activation free energies in the modelling dataset.

5.3 Most common substrates and nucleophiles The most common substrates and nucleophiles in the modelling dataset are given in Figure S19 and

Figure S20.

Figure S19. Most common substrates in the dataset with frequency indicated.

Page 48: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

27

Figure S20. Most common nucleophiles with frequency indicated.

5.4 Replicated reactions The reactions replicated in the full dataset are given in Table S12, of which the one carried out in TMS is

not included in the modelling dataset as solvent features are lacking for the TMS solvent.

Table S12. Replicated reactions in the full dataset.

Reaction Solvent Temperature (K) ΔG‡ (kcal/mol)

Acetonitrile 298 13.91, 14.46

Ethanol 298 19.80, 19.90

HMPT 298 17.29, 18.92

TMS 298 21.18a, 21.10a

a Not included in the modelling dataset as we don’t have solvent features for TMS.

5.5 Conditions Conditions are here defined in terms of both solvent and temperature. Figure S21 and Table S13 give the

number of reactions with more than one condition in the modelled dataset (446 reactions).

Page 49: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

28

Figure S21. Number of reactions with more than one condition in the modelling dataset.

Table S13. Number of reactions with more than one condition in the modelling dataset.

No. conditions 2 3 4 5 6 8 9

Counts 49 5 1 1 1 4 1

5.6 Reaction temperatures The distribution of reactions at different reaction temperatures for the modelling dataset are given in

Figure S22 and Table S14. Correlation analysis confirmed that exclusion of reaction temperature was

warranted (R: 0.64, ρ: 0.54 τ: 0.43) as it could possibly artificially inflate the performance of the model

(Figure S23). Reactions with higher barriers are often run at higher temperatures to be measured in

reasonable time.

Figure S22. Distribution of reaction with temperature in the modelling dataset. Note log scale.

Table S14. Reactions with a certain temperature in the modelling dataset.

Temperature 298 323 293 303 273 363 432.6 468.4 432.5 353 344

Counts 288 59 58 16 13 4 1 1 1 1 1

Page 50: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

29

Figure S23. Experimental activation energy as a function of reaction temperature.

6 Workflow database The database holding the calculations from the predict-SNAr workflow is a Python shelve database which

operates as a dictionary with keys and values. Each key is a unique identifier, holding a dictionary of the

calculation results which are described in Table S15. The database is housed in the folder

“machine_learning/2019-12-15” under the name “db”. It can be loaded with the following Python code:

import shelve d = shelve.open(“db”)

The workflow dataset includes a number of duplicate reactions that need to be removed before

modelling (see the notebooks referenced in Section 2).

Table S15. Keys for the workflow dataset dictionary.

Key Description of values Datatype

'solvent' Solvent SMILES string String 'clustering_energies' Energy correction from explicit solvent

model (a.u.) See table below

'agent' Flag for agent (acid/base) Bool/None 'temperature' Temperature (K) Float 'n_cpus' Number of CPUs Integer 'run_time' Run time (seconds) Float 'end_time' End time (YYYY-MM-DD HH:MM:SS) String 'flat_PES' Flag for flat PES where intermediate is

found but not elimination TS Bool

'intramolecular' Flag for intramolecular reaction Bool 'concerted' Flag for concerted reaction Bool 'descriptors' Descriptors See table below. 'reactive_atoms' Reactive atoms Dictionary, keys: string, values: integer 'symbols' Chemical symbols Dictionary, keys: string, values: list of

strings 'geometries' Geometries (Å) Dictionary, keys: string, values: list of

floats 'entropy_corr_qh_truhlar' Entropy terms with Truhlar quasi-

harmonic correction (a.u.) Dictionary, keys: string, values: float

'entropy_corr_qh_grimme' Entropy terms with Grimme quasi-harmonic correction (a.u.)

Dictionary, keys: string, values: float

'entropy_corr' Entropy terms (a.u.) Dictionary, keys: string, values: float 'enthalpy_corr' Enthalpy terms (a.u.) Dictionary, keys: string, values: float 'free_energies_qh_truhlar' Free energies with Truhlar quasi-

harmonic correction (a.u.) Dictionary, keys: string, values: float

'free_energies_qh_grimme' Free energies with Grimme quasi-harmonic correction (a.u.)

Dictionary, keys: string, values: float

Page 51: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

30

'free_energies' Free energies (a.u.) Dictionary, keys: string, values: float 'enthalpies' Enthalpies (a.u.) Dictionary, keys: string, values: float 'electronic_energies' Electronic energies (a.u.) Dictionary, keys: string, values: float 'inchi_key' InChIKey Dictionary, keys: string, values: string 'inchi' InChI Dictionary, keys: string, values: string 'smiles' SMILES String

The keys used in the sub-dictionaries are described in Table S16.

Table S16. Keys for sub-dictionaries in the workflow dataset.

Key Description

“substrate” Substrate “nucleophile” Nucleophile “product” Product “leaving_group” Leaving group or leaving atom “ts” Transition state “agent” Agent. Not implemented. “reaction” Related to the reaction, e.g., reaction SMILES “reaction_orig” Related to the reaction as originally input, e.g., reaction SMILES “solvent” Solvent

The keys for the descriptor sub-dictionary are given in Table S17.

Table S17. Keys for the descriptor sub-dictionary.

Key Description Value

'epn' Electrostatic potential at nuclei (VN, a.u.) Dictionary, keys: integer, values: float 'hirshfeld' Hirshfeld atomic charge Dictionary, keys: integer, values: float 'hirshfeld_plus' Hirshfeld atomic charge for cation Dictionary, keys: integer, values: float 'hirshfeld_minus' Hirshfeld atomic charge for anion Dictionary, keys: integer, values: float 'ddec6_charge' DDEC6 charge (q) Dictionary, keys: integer, values: float 'ddec6_bo' DDEC6 bond order (BO) Dictionary, keys: tuple of integers,

values: float 'v_av' Atom surface average of the electrostatic potential

(V̅s, kcal/mol) Dictionary, keys: integer, values: float

'vs_max' Atom surface maximum of the electrostatic potential (kcal/mol).

Dictionary, keys: integer, values: float

'vs_min' Atom surface minimum of the electrostatic potential (kcal/mol).

Dictionary, keys: integer, values: float

'es_min_b3lyp' Atomic surface minimum of the local electron attachment with the B3LYP functional (Es,min, eV).

Dictionary, keys: integer, values: float

'es_min_blyp' Atomic surface minimum of the local electron attachment with the BLYP functional (eV).

Dictionary, keys: integer, values: float

'is_min' Atomic surface minimum of the the average local ionization energy (Is,min, eV).

Dictionary, keys: integer, values: float

'sasa' Solvent accessible surface area (Å2) Dictionary, keys: integer, values: float 'sasa_ratio' Ratio of available solvent accessible surface area

(SASAr) Dictionary, keys: integer, values: float

'ip' Ionization potential (eV) Float 'ea' Ionization potential (eV) Float 'atom_p_int' Atomic dispersion potentials (kcal0.5 mol−0.5) Dictionary, keys: integer, values: float 'atom_p_int_area' Atomic dispersion potentials multiplied by atomic

area (kcal0.5 mol−0.5 Å2) Dictionary, keys: integer, values: float

'p_int' Average dispersion potential of entire molecule (kcal0.5 mol−0.5)

Float

'p_int_area' Average dispersion potential of entire molecule multiplied by its area (kcal0.5 mol−0.5 Å2)

Float

The subdictionary “clustering_energies” contains the solvent corrections from explicit solvation and are

explained in Table S18. It uses the keys 'xtb' and 'dft'.

Page 52: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

31

Table S18. Keys for the clustering energies sub-dictionary

Key Description

'clustering_energy' Solvation model correction (a.u.) 'clustering_energy_qh_grimme' Solvation model correction with Grimme correction (a.u.) 'clustering_energy_qh_truhlar' Solvation model correction with Truhlar correction (a.u.)

7 Analysis package A Python package, analyze_snar, was written for extraction of data from the workflow database as well

as machine learning routines and statistical functions. It is included with the manuscript and can be

installed with “pip install .”.

8 References 1. https://www.daylight.com/dayhtml/doc/theory/theory.smiles.html. 2. Reichardt, C. (2003). Solvents and Solvent Effects in Organic Chemistry 3rd ed. (Wiley-VCH,

Weinheim). 3. Lowe, D.M., Corbett, P.T., Murray-Rust, P., and Glen, R.C. (2011). Chemical Name to

Structure: OPSIN, an Open Source Solution. J. Chem. Inf. Model. 51, 739–753. 4. https://github.com/dan2097/opsin. 5. Diorazio, L.J., Hose, D.R.J., and Adlington, N.K. (2016). Toward a More Holistic Framework

for Solvent Selection. Org. Process Res. Dev. 20, 760–773. 6. RDKit: Open-source cheminformatics http://www.rdkit.org. 7. Ebejer, J.-P., Morris, G.M., and Deane, C.M. (2012). Freely Available Conformer Generation

Methods: How Good Are They? J. Chem. Inf. Model. 52, 1146–1158. 8. Riniker, S., and Landrum, G.A. (2015). Better Informed Distance Geometry: Using What We

Know To Improve Conformation Generation. J. Chem. Inf. Model. 55, 2562–2574. 9. Halgren, T.A. (1996). Merck molecular force field. I. Basis, form, scope, parameterization,

and performance of MMFF94. J. Comput. Chem. 17, 490–519. 10. Tosco, P., Stiefl, N., and Landrum, G. (2014). Bringing the MMFF force field to the RDKit:

implementation and validation. J. Cheminformatics 6, 37. 11. Rappe, A.K., Casewit, C.J., Colwell, K.S., Goddard, W.A., and Skiff, W.M. (1992). UFF, a full

periodic table force field for molecular mechanics and molecular dynamics simulations. J. Am. Chem. Soc. 114, 10024–10035.

12. Frisch, M.J., Trucks, G.W., Schlegel, H.B., Scuseria, G.E., Robb, M.A., Cheeseman, J.R., Scalmani, G., Barone, V., Petersson, G.A., Nakatsuji, H., et al. (2016). Gaussian 16 Rev. C.01.

13. Chai, J.-D., and Head-Gordon, M. (2008). Long-range corrected hybrid density functionals with damped atom-atom dispersion corrections. Phys. Chem. Chem. Phys. 10, 6615–6620.

14. Ditchfield, R., Hehre, W.J., and Pople, J.A. (1971). Self‐consistent molecular‐orbital methods. IX. An extended gaussian‐type basis for molecular‐orbital studies of organic molecules. J. Chem. Phys. 54, 724–728.

15. Clark, T., Chandrasekhar, J., Spitznagel, G.W., and Schleyer, P.V.R. (1983). Efficient diffuse function-augmented basis sets for anion calculations. III. The 3-21+G basis set for first-row elements, Li–F. J. Comput. Chem. 4, 294–301.

16. Chéron, N., Jacquemin, D., and Fleurat-Lessard, P. (2012). A qualitative failure of B3LYP for textbook organic reactions. Phys. Chem. Chem. Phys. 14, 7170–7175.

17. Grimme, S. (2012). Supramolecular Binding Thermodynamics by Dispersion-Corrected Density Functional Theory. Chem. - Eur. J. 18, 9955–9964.

Page 53: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

32

18. Luchini, G., Alegre-Requena, J.V., Funes-Ardoiz, I., and Paton, R.S. (2020). GoodVibes: automated thermochemistry for heterogeneous computational chemistry data. F1000Research 9, 291.

19. https://github.com/bobbypaton/GoodVibes. 20. Krishnan, R., Binkley, J.S., Seeger, R., and Pople, J.A. (1980). Self‐consistent molecular orbital

methods. XX. A basis set for correlated wave functions. J. Chem. Phys. 72, 650–654. 21. Marenich, A.V., Cramer, C.J., and Truhlar, D.G. (2009). Universal Solvation Model Based on

Solute Electron Density and on a Continuum Model of the Solvent Defined by the Bulk Dielectric Constant and Atomic Surface Tensions. J. Phys. Chem. B 113, 6378–6396.

22. Bannwarth, C., Ehlert, S., and Grimme, S. (2019). GFN2-xTB—An Accurate and Broadly Parametrized Self-Consistent Tight-Binding Quantum Chemical Method with Multipole Electrostatics and Density-Dependent Dispersion Contributions. J. Chem. Theory Comput. 15, 1652–1671.

23. https://github.com/grimme-lab/xtb. 24. Pliego, J.R., and Riveros, J.M. (2001). The Cluster−Continuum Model for the Calculation of

the Solvation Free Energy of Ionic Species. J. Phys. Chem. A 105, 7241–7247. 25. Pliego Jr, J.R., and Riveros, J.M. (2019). Hybrid discrete-continuum solvation methods. Wiley

Interdiscip. Rev. Comput. Mol. Sci., e1440. 26. https://github.com/grimme-lab/crest. 27. Plata, R.E., and Singleton, D.A. (2015). A Case Study of the Mechanism of Alcohol-Mediated

Morita Baylis–Hillman Reactions. The Importance of Experimental Observations. J. Am. Chem. Soc. 137, 3811–3826.

28. Terrier, F. (2013). Modern Nucleophilic Aromatic Substitution (Wiley-VCH). 29. Weigend, F., and Ahlrichs, R. (2005). Balanced basis sets of split valence, triple zeta valence

and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys. Chem. Chem. Phys. 7, 3297–3305.

30. Rappoport, D., and Furche, F. (2010). Property-optimized Gaussian basis sets for molecular response calculations. J. Chem. Phys. 133, 134105.

31. Manuscript in preparation. 32. Shrake, A., and Rupley, J.A. (1973). Environment and exposure to solvent of protein atoms.

Lysozyme and insulin. J. Mol. Biol. 79, 351–371. 33. Haynes, W.M., and Haynes, W.M. (2014). CRC Handbook of Chemistry and Physics 95th ed. 34. Rahm, M., Hoffmann, R., and Ashcroft, N.W. (2016). Atomic and Ionic Radii of Elements 1–

96. Chem. - Eur. J. 22, 14625–14632. 35. Grimme, S., Antony, J., Ehrlich, S., and Krieg, H. (2010). A consistent and accurate ab initio

parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 132, 154104.

36. Brinck, T., Carlqvist, P., and Stenlid, J.H. (2016). Local Electron Attachment Energy and Its Use for Predicting Nucleophilic Reactions and Halogen Bonding. J. Phys. Chem. A 120, 10023–10032.

37. Sjoberg, P., Murray, J.S., Brinck, T., and Politzer, P. (1990). Average local ionization energies on the molecular surfaces of aromatic systems as guides to chemical reactivity. Can. J. Chem. 68, 1440–1443.

38. Brinck, T. HS95: Molecular surface property program. 39. Manz, T.A., and Gabaldon Limas, N. (2017). Chargemol program for performing DDEC

analysis. 40. https://sourceforge.net/projects/ddec/.

Page 54: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

33

41. 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.

42. Parr, R.G., Szentpály, L. v., and Liu, S. (1999). Electrophilicity Index. J. Am. Chem. Soc. 121, 1922–1924.

43. Oller, J., Pérez, P., Ayers, P.W., and Vöhringer-Martinez, E. (2018). Global and local reactivity descriptors based on quadratic and linear energy models for α,β-unsaturated organic compounds. Int. J. Quantum Chem. 118, e25706.

44. Hirshfeld, F.L. (1977). Bonded-atom fragments for describing molecular charge densities. Theor. Chim. Acta 44, 129–138.

45. BIOVIA Pipeline Pilot (2018). (Dassault Systèmes). 46. Nugmanov, R.I., Mukhametgaleev, R.N., Akhmetshin, T., Gimadiev, T.R., Afonina, V.A.,

Madzhidov, T.I., and Varnek, A. (2019). CGRtools: Python Library for Molecule, Reaction, and Condensed Graph of Reaction Processing. J. Chem. Inf. Model. 59, 2516–2521.

47. https://github.com/cimm-kzn/CGRtools. 48. https://github.com/cimm-kzn/CIMtools. 49. Philippe Schwaller, Daniel Probst, Alain C. Vaucher, Vishnu H Nair, Teodoro Laino, and Jean-

Louis Reymond (2019). Data-Driven Chemical Reaction Classification, Fingerprinting and Clustering using Attention-Based Neural Networks. ChemRxiv.

50. https://github.com/rxn4chemistry/rxnfp/. 51. Hjorth Larsen, A., Jørgen Mortensen, J., Blomqvist, J., Castelli, I.E., Christensen, R., Dułak,

M., Friis, J., Groves, M.N., Hammer, B., Hargus, C., et al. (2017). The atomic simulation environment—a Python library for working with atoms. J. Phys. Condens. Matter 29, 273002.

52. O’boyle, N.M., Tenderholt, A.L., and Langner, K.M. (2008). cclib: A library for package-independent computational chemistry algorithms. J. Comput. Chem. 29, 839–845.

53. Pedregosa, F., Varoquaux, G., Gramfort, A., Michel, V., Thirion, B., Grisel, O., Blondel, M., Prettenhofer, P., Weiss, R., Dubourg, V., et al. (2011). Scikit-learn: Machine Learning in Python. J. Mach. Learn. Res. 12, 2825–2830.

54. McKinney, W. (2010). Data Structures for Statistical Computing in Python. In Proceedings of the 9th Python in Science Conference, S. van der Walt and J. Millman, eds., pp. 56–61.

55. Hunter, J.D. (2007). Matplotlib: A 2D Graphics Environment. Comput. Sci. Eng. 9, 90–95. 56. https://github.com/mwaskom/seaborn. 57. Pearson, K. (1895). VII. Note on regression and inheritance in the case of two parents. Proc.

R. Soc. Lond. 58, 240–242. 58. James, G., Witten, D., Hastie, T., and Tibshirani, R. (2013). An introduction to statistical

learning: with applications in R (Springer). 59. Tsamardinos, I., Greasidou, E., and Borboudakis, G. (2018). Bootstrapping the out-of-sample

predictions for efficient and accurate cross-validation. Mach. Learn. 107, 1895–1922. 60. Borboudakis, G., Stergiannakos, T., Frysali, M., Klontzas, E., Tsamardinos, I., and Froudakis,

G.E. (2017). Chemically intuited, large-scale screening of MOFs by machine learning techniques. Npj Comput. Mater. 3, 40.

61. Cawley, G.C., and Talbot, N.L. (2010). On over-fitting in model selection and subsequent selection bias in performance evaluation. J. Mach. Learn. Res. 11, 2079–2107.

62. Boser, B.E., Guyon, I.M., and Vapnik, V.N. (1992). A Training Algorithm for Optimal Margin Classifiers. In Proceedings of the Fifth Annual Workshop on Computational Learning Theory COLT ’92. (Association for Computing Machinery), pp. 144–152.

Page 55: Machine Learning Meets Mechanistic Modelling for Accurate ... · E Ar), single-task deep learning models trained for selectivity prediction of bromination, chlorination, nitration

34

63. Breiman, L. (2001). Random Forests. Mach. Learn. 45, 5–32. 64. Friedman, J.H. (2001). Greedy function approximation: A gradient boosting

machine. Ann Stat. 29, 1189–1232. 65. Rasmussen, C.E., and Williams, C.K.I. (2006). Gaussian processes for machine learning (MIT

Press). 66. Neal, R.M. (1996). Bayesian learning for neural networks (Springer). 67. Wipf, D.P., and Nagarajan, S.S. (2008). A New View of Automatic Relevance Determination.

In Advances in Neural Information Processing Systems 20, J. C. Platt, D. Koller, Y. Singer, and S. T. Roweis, eds. (Curran Associates, Inc.), pp. 1625–1632.

68. Wold, S., Sjöström, M., and Eriksson, L. (2001). PLS-regression: a basic tool of chemometrics. Chemom. Intell. Lab. Syst. 58, 109–130.

69. Altman, N.S. (1992). An Introduction to Kernel and Nearest-Neighbor Nonparametric Regression. Am. Stat. 46, 175–185.

70. Pearson, K. (1901). LIII. On lines and planes of closest fit to systems of points in space. Lond. Edinb. Dublin Philos. Mag. J. Sci. 2, 559–572.

71. McInnes, L., Healy, J., and Melville, J. (2018). UMAP: Uniform Manifold Approximation and Projection for Dimension Reduction. arXiv:1802.03426.

72. McInnes, L., Healy, J., Saul, N., and Großberger, L. (2018). UMAP: Uniform Manifold Approximation and Projection. J. Open Source Softw. 3, 861.

73. Probst, D., and Reymond, J.-L. (2020). Visualization of very large high-dimensional data sets as minimum spanning trees. J. Cheminformatics 12, 12.

74. https://github.com/reymond-group/tmap. 75. Probst, D., and Reymond, J.-L. (2018). FUn: a framework for interactive visualizations of

large, high-dimensional datasets on the web. Bioinformatics 34, 1433–1435. 76. https://github.com/reymond-group/faerun. 77. SciPy 1.0 Contributors, Virtanen, P., Gommers, R., Oliphant, T.E., Haberland, M., Reddy, T.,

Cournapeau, D., Burovski, E., Peterson, P., Weckesser, W., et al. (2020). SciPy 1.0: fundamental algorithms for scientific computing in Python. Nat. Methods 17, 261–272.

78. Spearman, C. (1904). The Proof and Measurement of Association between Two Things. Am. J. Psychol. 15, 72.

79. Raschka, S. (2018). MLxtend: Providing machine learning and data science utilities and extensions to Python’s scientific computing stack. J. Open Source Softw. 3.

80. Rücker, C., Rücker, G., and Meringer, M. (2007). y-Randomization and Its Variants in QSPR/QSAR. J. Chem. Inf. Model. 47, 2345–2357.

81. Ferreira, M.M.C. (2017). Multivariate QSAR. In Encyclopedia of Physical Organic Chemistry (American Cancer Society), pp. 1–38.

82. Schneider, N., Stiefl, N., and Landrum, G.A. (2016). What’s What: The (Nearly) Definitive Guide to Reaction Role Assignment. J. Chem. Inf. Model. 56, 2336–2346.

83. NameRxn (2019). (NextMove Software Limited).