Skip to main content

Critical assessment of synthetic accessibility scores in computer-assisted synthesis planning

Abstract

Modern computer-assisted synthesis planning tools provide strong support for this problem. However, they are still limited by computational complexity. This limitation may be overcome by scoring the synthetic accessibility as a pre-retrosynthesis heuristic. A wide range of machine learning scoring approaches is available, however, their applicability and correctness were studied to a limited extent. Moreover, there is a lack of critical assessment of synthetic accessibility scores with common test conditions.In the present work, we assess if synthetic accessibility scores can reliably predict the outcomes of retrosynthesis planning. Using a specially prepared compounds database, we examine the outcomes of the retrosynthetic tool AiZynthFinder. We test whether synthetic accessibility scores: SAscore, SYBA, SCScore, and RAscore accurately predict the results of retrosynthesis planning. Furthermore, we investigate if synthetic accessibility scores can speed up retrosynthesis planning by better prioritizing explored partial synthetic routes and thus reducing the size of the search space. For that purpose, we analyze the AiZynthFinder partial solutions search trees, their structure, and complexity parameters, such as the number of nodes, or treewidth.We confirm that synthetic accessibility scores in most cases well discriminate feasible molecules from infeasible ones and can be potential boosters of retrosynthesis planning tools. Moreover, we show the current challenges of designing computer-assisted synthesis planning tools. We conclude that hybrid machine learning and human intuition-based synthetic accessibility scores can efficiently boost the effectiveness of computer-assisted retrosynthesis planning, however, they need to be carefully crafted for retrosynthesis planning algorithms.The source code of this work is publicly available at https://github.com/grzsko/ASAP.

Introduction

Table 1 Comparison of analyzed synthetic accessibility scores

The present era of machine learning (ML) and deep learning (DL) techniques and high computing power provides solutions to problems previously treated as untractable. One of them is computer-assisted synthesis planning (CASP) which consists of two tasks: reactions forward planning and retrosynthesis. The former is predicting the outcomes of reaction for given reactants. The latter is a method of planning the synthesis scheme of chemical compounds from simple precursors available in stock, to synthesized intermediates, and the target molecule. Synthesis planning remained a laborious, manual task until the 1960s when Corey [1] formalized the idea of CASP and then implemented it in LHASA [2] software. Over the years, new solutions were developed that automated subsequent planning elements, required less human intervention, and increased the speed and accuracy of algorithms [3,4,5]. Over the last decade, several modern, ML-based CASP tools were independently developed: from closed vendor software, e.g. Synthia (previously Chematica) [6, 7], to the closed source with the available interface, e.g. IBM RXN [8], and open-source ones, e.g. LillyMol [9], AiZynthFinder [10,11,12], ASKCOS Tree-builder [13], AutoSynRoute [14]. Currently, a standard CASP tool [15] consists of three modules: (i) the database of reaction templates and rules on how to apply them to analyzed molecules, (ii) algorithms searching for possible synthetic routes, (iii) a database of in-stock molecules. The aforementioned tools differ significantly in the design of every module. For example, the database of reaction templates may be manually encoded with a rule-based algorithm for reaction prediction, e.g. Synthia. It may be also automatically extracted and reactions may be predicted with a neural network, e.g. LillyMol, AiZynthFinder, and ASKCOS Tree-builder. Finally, reactions may be predicted using a template-free seq2seq algorithm [16] known from natural language processing as implemented in IBM RXN.

Besides CASP tools’ strengths, their key bottleneck is computational complexity. During retrosynthesis planning runtime, potentially exponential in size search space of solution candidates (partial synthetic routes) must be traversed. It makes CASP tools non-applicable when numerous molecules need to be immediately checked for synthesizability. One example is a virtual screening (VS) method known in computer-assisted drug design (CADD). During VS, even billions of compound candidates are evaluated for desired properties; thus, searching for a synthetic route for each of these candidates is computationally intractable.

This limitation may be overcome by scoring the synthetic accessibility, i.e. by predicting how the molecule of a given structure is synthesizable. Previously, synthetic accessibility scores were based on single molecular properties selected manually by experts [17,18,19,20]. With the emergence of ML and DL methods, new scores were designed. They can be divided into structure-based and reaction-based approaches. Structure-based approaches evaluate the feasibility of molecular structure, e.g. SAscore [21], SYBA [22], GASA [23]. Reaction-based approaches predict the synthetic accessibility by capturing the similarity of synthetic routes deposited in reaction databases, e.g. SCScore [24], RAscore [25], CMPNN [26], or RetroGNN [27].

Although the majority of these scores are publicly available and documented, their applicability as a pre-retrosynthesis heuristic is known to a limited extent. Moreover, there is a lack of critical assessment of synthetic accessibility scores on the standardized dataset with common test conditions.

In the present work, we assess if synthetic accessibility scores can reliably predict outcomes of retrosynthesis planning. We also analyze if synthetic accessibility scores can speed up the retrosynthesis planning by reducing the size of the search space. Specifically, we analyze the outcomes and runtime of the retrosynthetic tool AiZynthFinder on a specially prepared compounds database. We assess if four scores: SAscore, SCScore, RAscore, and SYBA (cf. Table 1) properly predict the results of retrosynthesis planning and the search complexity. To do this, we analyze the AiZynthFinder partial solutions search trees. Moreover, by in-depth analysis of these search trees, we assess if synthetic accessibility scores can speed up retrosynthesis planning by better prioritizing partial synthetic routes.

To the best of the authors’ knowledge, it is the first of this kind of assessment. Although benchmarks are available in cheminformatics, they focus on the outputs of the CASP tools [28] or on synthetic accessibility scores alone [29, 30]. This assessment is easily reproducible and is designed as a framework for evaluating and comparing novel synthetic accessibility scores. Its source code with usage instructions is publicly available at https://github.com/grzsko/ASAP.

Methods

Fig. 1
figure 1

AiZynthFinder search tree nodes classification. Nodes are classified as: internal (I), solved (S), and not solved (NS). Internal nodes have a non-empty list of expandable molecules, but their depth is below a predefined depth. Solved nodes are leaves with all molecules in the in-stock list. A leaf marked as not solved means that it contains at least one expandable molecule and its depth exceeds a predefined threshold. Because we aim to discriminate promising nodes from non-promising ones as early as possible, we define a not solved node as all nodes that have no path to a solved leaf. In the majority of cases, we focus on roots of subtrees of not solved nodes

Analyzed synthetic accessibility scores

SAscore

SAscore [21] is designed as a synthetic accessibility score of drug-like molecules for virtual screening exploration. It is calculated as a sum of fragment scores and complexity penalty. Fragment score is based on statistics of the frequency of Extended Connectivity Fingerprints of diameter 4 (ECFP4) [31] fragments from Pipeline Pilot [32] on almost one million molecules obtained from the PubChem database [33]. ECFP is a method of creating a numeric representation of a chemical structure by traversing it, enumerating atoms, and hashing their representation. The aim of the fragment score is to capture if fragments observed previously in the database are present in the analyzed molecule. The complexity penalty aims to capture if a molecule does not contain too many complex structures to be synthesized. It incorporates among others number of aromatic rings, stereocenters, macrocycles, or the size of the molecule. SAscore achieves values from 1 (easy to synthesize) to 10 (hard to synthesize). It is publicly available in RDKit package [34].

SYBA

The idea of the SYBA score is to train a model on comprehensive representations of both existing, easy-to-synthesize compounds as well as non-existing, hard-to-synthesize compounds. The former set was randomized from the ZINC15 database and the latter set was created from an easy-to-synthesize one using Nonpher tool [35] by the iterative perturbing structure of the input molecules (adding/removing of atom or bond) up to a predefined complexity threshold. SYBA is a Bernoulli naïve Bayes classifier trained on both sets. Its implementation is available as a Conda package or at https://github.com/lich-uct/syba.

SCScore

SCScore is a score for assessing the molecular complexity expressed as the expected number of reaction steps required to produce a target. This score was trained using neural networks [36] on the set of 12 million reactions obtained from the Reaxys database [37]. Molecules for this score are represented as 1024-bit Morgan fingerprints of radius 2 [38] which are generally similar to ECFP4. It achieves values from 1 (simple molecule) to 5 (complex molecule). This score was used as precursor prioritizer in ASKCOS Tree-builder tool [13] and is publicly available in GitHub repository https://github.com/connorcoley/scscore.

RAscore

RAscore is designed as a retrosynthetic accessibility score, i.e. score for fast prescreening molecules for the AiZynthFinder tool. It was trained on over 200000 molecules from ChEMBL [39]. For every molecule, a synthesis route was generated using AiZynthFinder to assess if the molecule is synthesizable. Two models were trained on these outcomes: neural network [36] and gradient boosting machine [40]. RAscore implementation is publicly available at https://github.com/reymond-group/RAscore.

Analyzed CASP tool

AiZynthFinder is an algorithm for computational synthesis planning. It utilizes the Monte Carlo tree search (MCTS) algorithm [41, 42], which is used for searching the tree of possible partial solutions to the analyzed problem. Here, solutions correspond to synthetic routes of the target molecule. Single MCTS round consists of 4 steps [43]: (1) selection of random leaf node, (2) expansion during which new nodes from leaf are created, (3) rollout, i.e. search simulation from new node till the complete solution or a partial solution exceeding a predefined depth, (4) backpropagation during which nodes are actualized after rollout. The node containing a partial solution is represented by (i) its depth, (ii) the set of in-stock molecules, and (iii) the set of expandable molecules which need to be further transformed into simpler, buyable molecules. Here, the depth of the node is defined as the maximal number of transformations that each of its molecules has to undergo to the target. A leaf node represents a complete solution if it does not need to be expanded, i.e. its list of expandable molecules is empty and its depth does not exceed a predefined threshold. Otherwise, a leaf node represents an infeasible partial solution with a depth exceeding a threshold i.e. it corresponds to the too long synthetic route. The root node of the search tree contains a single expandable molecule representing the target compound. Nodes are connected with directed edges representing a reaction whose product is a single expandable molecule. Leaf selection is made by recursively traversing a search tree starting from the root by selecting children of maximum upper confidence bound (UCB) which expresses current node exploitation and how it is promising:

$$\begin{aligned} UCB = \frac{Q}{N_p} + U. \end{aligned}$$
(1)

U describes how the node was already explored, i.e.

$$\begin{aligned} U = 1.4\cdot \sqrt{2\cdot \frac{\ln N_{-1}}{N_p}}, \end{aligned}$$

where \(N_p\) is the number of times the child node has been visited, and \(N_{-1}\) is the number of times the parent node has been visited. Q describes how the node is promising, i.e. it is a sum of rewards from previous backpropagations. A single reward equals:

$$\begin{aligned} 0.95 \cdot \frac{M_{\text {s}}}{M} + 0.05\ \frac{1}{1+\exp (m - 4)}, \end{aligned}$$
(2)

where M is the number of molecules in the node, \(M_{\text {s}}\) is the number of solved molecules and m is the maximum number of transformations that every molecule have to undergo to become the root. A reward assesses how molecules of a given node are already expanded and how many steps are used. Nodes are expanded using a neural network applying reaction templates on expandable molecules in the node. Reactions are chosen so that the UCB of the product is maximized.

Evaluation of synthesis planning and scores

Dataset

We prepared a database of 49 compounds. Their detailed list is available in Additional File 1. The majority of these compounds are drugs or plant metabolites, of which 44 have documented synthesis. Molecules in our database were collected to represent various synthesis complexity, starting from easily synthesizable ones such as acetylsalicylic acid, to compounds of known synthesis but the more complex structure, such as morphine, compounds of known low yielding synthesis, such as isocorydine, and not known to be synthesizable. On the other hand, the molecules were collected to represent several examples of high demand for synthesizability, such as drugs, plant metabolites, human metabolites, etc. All compounds have their structure encoded in SMILES notation [44] from PubChem [33] with incorporating stereo orientation. Molecules from this database were further input dataset of AiZynthFinder tool and synthetic accessibility scores for their analysis.

Fig. 2
figure 2

Analyzed nodes configurations Panel A: We checked if two nodes, internal and not solved which have the same internal parent can be discriminated by synthetic accessibility scores. Panel B: We checked also if synthetic accessibility scores can discriminate internal parents from their not solved children.

Analysis of the search trees

In the first analysis, we assessed if synthetic accessibility scores can model and predict outcomes of retrosynthesis planning. To express the complexity of retrosynthesis planning, we analyzed the search trees of AiZynthFinder runtime for molecules from our database. For these trees, we calculated statistics, such as the number of nodes, treewidth, and the number of leaf nodes that are not solved. We omitted to analyze tree depth because AiZynthFinder has strict limits for the depth of the search tree and the results would be uninformative.

Moreover, we checked if synthetic accessibility scores can act as nodes’ prioritization heuristics. To this end, we classified all nodes into three groups: solved, not solved, and internal (cf. Fig. 1). Solved nodes correspond to complete solutions, i. e. all their molecules are available in stock. Not solved nodes correspond to partial solutions of an infeasible synthetic route. We define not solved nodes as nodes for which there is no path leading to a solved node. The rest of the nodes are internal, i.e. nodes having a path to the solved leaf node. They correspond to these partial solutions which eventually lead to a complete solution. For such nodes definition, if the root of a tree is not solved then the algorithm has not found any feasible synthetic route for a given target molecule. We express a score value of a node as one of the statistics (maximum, minimum, arithmetic mean) over all molecules in the node. For making calculations comparable, all scores were transformed so that they achieve values from the range [0, 1] with 0 corresponding to an infeasible (non-synthesizable) molecule and 1 corresponding to a feasible (easily synthesizable) molecule. To check if synthetic accessibility scores properly prioritize nodes, we analyzed if synthetic accessibility scores discriminate internal nodes from not solved ones. Firstly, we considered these pairs connected with a single reaction step. We analyzed two configurations: (i) siblings nodes internal and not solved with internal parent and (ii) internal parent from not solved child (cf. Fig. 2). Secondly, we checked if synthetic accessibility scores correctly discriminate internal nodes from not solved ones in general.

Finally, we checked if modified leaf selection, which incorporates nodes’ synthetic accessibility scores, may speed up retrosynthesis planning. To this point, we modified UCB (Eq. (1)) by substituting a fraction of a reward with one of the synthetic accessibility scores. Specifically, a reward (Eq. (2)) was replaced with the value:

$$\begin{aligned} c\cdot \mathcal{S}\mathcal{A} + (0.95-c) \cdot \frac{M_{\text {s}}}{M} + 0.05\ \frac{1}{1+\exp (m - 4)}, \end{aligned}$$
(3)

where c is a replaced fraction of reward (\(\frac{1}{4}\cdot 0.95\), \(\frac{2}{4}\cdot 0.95\), \(\frac{3}{4}\cdot 0.95\)) and \(\mathcal{S}\mathcal{A}\) is one of appriopriately transformed synthetic accessibility scores.

Results and discussion

Fig. 3
figure 3

ROC curve for synthetic accessibility scores prediction of AiZynthFinder outcomes. Dots mark the best score threshold. AUCs for curves are listed in Table 2

Table 2 Comparison of analyzed synthetic accessibility scores in predicting the AiZynthFinder outcomes
Fig. 4
figure 4

Heatmap of correlation between synthetic accessibility scores and complexity search tree parameters. Colors and white labels indicate the value of the Spearman correlation, black labels indicate the p-value of the correlation test.

Fig. 5
figure 5

Heatmaps of t-test p-values for hypothesis whether synthetic accessibility scores discriminate node types. Panel A: For internal and not solved siblings node pairs and solved and not solved node pairs if their scaled score differences are greater than 0. Panel B: For internal parent and not solved child node pairs if their scaled score differences are greater than 0. Here, discrimination between solved and not solved is not applicable.

For all compounds from our database, we performed retrosynthesis planning using AiZynthFinder with default parameters. AiZynthFinder found a synthetic route for 22 compounds. For all found synthetic routes, 20 of them are known (precision 0.91), and for all known synthetic routes, 20 of them are found (sensitivity 0.45).

We assessed if synthetic accessibility scores correctly predicted the results of retrosynthetic planning. To find the optimal score thresholds that discriminate synthesizable target molecules from non-synthesizable ones, we analyzed a receiver operating curve (ROC), cf. Fig. 3. It allows for finding the best balance between the sensitivity and specificity of the classifier. For every score and its optimal threshold, we computed the prediction accuracy of AiZynthFinder’s outcomes. We also measured the quality of scores by calculating the area under the ROC curve (AUC) which describes the probability that a score ranks a randomly chosen synthesizable molecule better than a randomly chosen non-synthesizable molecule. Results are depicted in Table 2. For both AUC and accuracy, SAscore and RAscore achieves high results (AUC and accuracy were both over 0.81). On the contrary, for both SCScore and SYBA, the results are worse by about 20 percentage points. RAscore’s good result is not surprising, because it was trained on the outcomes of the AiZynthFinder algorithm. This, combined with the low sensitivity of AiZynthFinder, allows us to claim that RAscore is a precise heuristic of AiZynthFinder outcomes, but not necessarily a synthetic accessibility score in general. The results of the SAscore may seem surprising. It is a slightly different score from the rest because it is not a standard ML model. It is designed as a combination of scores and penalties derived by experts from the presence of structural fragments in the PubChem database. From this, we infer that in retrosynthesis, human intuition and the power of the human mind still play an important role in planning a synthesis route especially in noticing the irregularities in the general synthesis rules. On the opposite, ML models are prone to imperfections, imbalance, bias, or gaps in training data. This lies in line with recent studies indicating ML limits in cheminformatics, for example for reaction yield prediction [45], for CADD [46], or for graph-based DL models for drug representation [47].

Fig. 6
figure 6

ROC curves for discrimination of internal and not solved nodes by appropriately scaled synthetic accessibility scores. AUCs are depicted in Fig. 7

Fig. 7
figure 7

Heatmap of AUC of discrimination between internal and not solved nodes (left) and solved and not solved nodes (right)

We checked also if synthetic accessibility scores can model the complexity of the retrosynthesis planning. We computed a Spearman rank correlation [48] between scores of target compounds and their search tree complexity parameters, such as treewidth, number of nodes, and number of not solved leaf nodes. Results are available in Fig. 4. All of RAscore, SAscore, and SCScore with at least one node aggregating statistic correlate negatively with all complexity parameters with significance below 0.04. On the contrary, SYBA does not correlate with any of the complexity parameters. Analogously as earlier, RAscore and SAscore performed best, the strongest negative correlation was observed between these two scores and the number of nodes.

As a next step, we checked if scores can be a good heuristic for prioritizing nodes corresponding to partial solutions. Well-prioritized nodes would preferably select routes that are more promising for further search and boost the efficiency of retrosynthesis planning. To this end, we checked if synthetic accessibility scores can detect potentially infeasible partial synthesis routes. We assessed this by taking all pairs of internal and not solved siblings nodes and checking if the average score of internal nodes is greater than the score of not solved nodes (cf. Fig. 2A). We used a one-sample t-test [49] for score differences of node pairs. The alternative hypothesis was that the mean of the pair differences distribution is greater than 0. We checked also if incorporating in-stock set molecules would not bias the node statistics. Thus, we repeated the same test on node scores incorporating only expandable molecules. Results are depicted in Fig. 5A. Practically, all scores with at least one aggregating statistic can correctly discriminate internal nodes from not solved and solved nodes from not solved. Omitting the set of in-stock molecules did not change the results.

We repeated the same analysis for pairs of the internal parent node and not solved child (cf. Fig. 2B). Contrary to previous results, only SAscore can significantly discriminate the parent internal node from its not solved child (cf. Fig. 5B).

Moreover, we checked in-depth if synthetic accessibility scores can correctly discriminate internal nodes from not solved ones and solved nodes from not solved ones. We collected all internal, solved, and not solved nodes. To find a threshold properly discriminating nodes, we analyzed ROC curves of synthetic accessibility scores, cf. Fig. 6 and Additional file 2: Figure S1. AUCs are depicted in Fig. 7. Practically, all scores except SYBA correctly discriminate internal nodes from not solved and solved from not solved. Note that for each of the rest of the scores, only the mean and minimum aggregating functions are efficient. It is because minimum detects the presence of non-synthesizable outliers while maximum reports the best synthesizable molecules. Analogously as earlier, SAscore achieved the best results and RAscore was slightly worse. The rest of the scores were considerably worse.

Finally, we analyzed if directly replacing a fraction of the reward with an appropriately scaled synthetic accessibility score may boost the retrosynthesis planning as in Eq. 3. If so then nodes during leaves selection would be better prioritized by UCB and thus computation time decreased. This replacement, however, did not significantly improve any parameter of search tree complexity (cf. Additional file 2: Tables S1–S4). It may be caused by undermined reward fraction in UCB formula (1) or high fitting of search algorithm design to its internal scorings. It should be noticed that the reward function is only used in the backpropagation phase and is calculated by the MTCS procedure in the leaf nodes to update the statistics of the search tree. On the other hand, the molecules at leaf nodes score highly on the synthesizability scale, as they are typically small and often purchasable. Therefore, the alternative approach to guide the selection of nodes would be adding the synthesizability scores to the UCB statistics calculated in the internal nodes. Such modification is worth implementing and we plan to incorporate it in further work.

Conclusions

In the present work, we analyzed if synthetic accessibility scores can effectively boost the retrosynthesis process. Our analyses consisted of checking if synthetic accessibility scores correctly model retrosynthesis planning outcomes and effectively discriminate feasible partial synthetic routes from infeasible ones. We confirmed that synthetic accessibility scores can in the majority of cases well discriminate feasible molecules from infeasible ones and can be potential boosters of retrosynthesis planning tools.

Today, the big-data era requires retrosynthesis planning tools to be a fast and accurate replacement for laborious, human-mind-based manual work. We show, however, that designing retrosynthesis planning algorithms is still a challenging task and require constant improvement for faster runtime and more accurate results. For example, replacing a fraction of UCB failed to improve AiZynthFinder accuracy which suggests that synthetic accessibility scores need to be carefully crafted for the target tool.

Moreover, high, outlying SAscore results suggest that currently, pure ML techniques still do not replace completely a human mind in the retrosynthesis planning process. This implies that the accuracy of scores, although increasing, is still limited. This results in a constant need for improving the quality of training datasets, because ML models may overfit to specific properties of training datasets that appeared to be unbalanced or biased. Also, there should be constant pressure for better model design. We conclude that hybrid ML and human intuition-based synthetic accessibility scores with carefully crafted retrosynthesis planning algorithms can still efficiently boost the effectiveness of computer-assisted retrosynthesis planning. These tools may help for both finding synthetic routes of newly designed compounds as well as recognizing what is still unknown in chemistry.

Availability of data and materials

Project name: ASAP - Critical Assessment of Synthetic Accessibility scores in computer-assisted synthesis Planning, Project home page: https://github.com/grzsko/ASAP, Operating system(s): Linux or macOS, Programming language: Python 3, Other requirements: Conda package management system, License: MIT, Any restrictions to use by non-academics: none. The molecule dataset being the input of AiZynthFinder is available in Additional file 1.

Abbreviations

AUC:

Area under the ROC curve

CADD:

Computer-assisted drug design

CASP:

Computer-assisted synthesis planning

DL:

Deep learning

ML:

Machine learning

MCTS:

Monte Carlo tree search

ROC:

Receiver operating curve

UCB:

Upper confidence bound

VS:

Virtual screening

References

  1. Corey EJ (1967) General methods for the construction of complex molecules. Pure Appl Chem 14(1):19–38

    Article  CAS  Google Scholar 

  2. Corey EJ, Cramer RD, Howe WJ (1972) Computer-assisted synthetic analysis for complex molecules. Methods and procedures for machine generation of synthetic intermediates. J Am Chem Soc 94(2):440–459

    Article  CAS  Google Scholar 

  3. Hanessian S, Franco J, Larouche B (1990) The psychobiological basis of heuristic synthesis planning - man, machine and the chiron approach. Pure Appl Chem 62(10):1887–1910

    Article  CAS  Google Scholar 

  4. Ihlenfeldt W-D, Gasteiger J (1996) Computer-assisted planning of organic syntheses: the second generation of programs. Angew Chem Int Ed Engl 34(23–24):2613–2633

    Article  Google Scholar 

  5. Ugi I, Bauer J, Bley K, Dengler A, Dietz A, Fontain E, Gruber B, Herges R, Knauer M, Reitsam K, Stein N (1993) Computer-assisted solution of chemical problems-the historical development and the present state of the art of a new discipline of chemistry. Angew Chem Int Ed Engl 32(2):201–227

    Article  Google Scholar 

  6. Szymkuć S, Gajewska EP, Klucznik T, Molga K, Dittwald P, Startek M, Bajczyk M, Grzybowski BA (2016) Computer-assisted synthetic planning: the end of the beginning. Angew Chem Int Ed 55(20):5904–5937

    Article  Google Scholar 

  7. Klucznik T, Mikulak-Klucznik B, McCormack MP, Lima H, Szymkuć S, Bhowmick M, Molga K, Zhou Y, Rickershauser L, Gajewska EP, Toutchkine A, Dittwald P, Startek MP, Kirkovits GJ, Roszak R, Adamski A, Sieredzińska B, Mrksich M, Trice SLJ, Grzybowski BA (2018) Efficient syntheses of diverse, medicinally relevant targets planned by computer and executed in the laboratory. Chem 4(3):522–532

    Article  CAS  Google Scholar 

  8. Schwaller P, Petraglia R, Zullo V, Nair VH, Haeuselmann RA, Pisoni R, Bekas C, Iuliano A, Laino T (2020) Predicting retrosynthetic pathways using transformer-based models and a hyper-graph exploration strategy. Chem Sci 11(12):3316–3325

    Article  CAS  Google Scholar 

  9. Watson IA, Wang J, Nicolaou CA (2019) A retrosynthetic analysis algorithm implementation. J Cheminformatics 11:1

  10. Genheden S, Thakkar A, Chadimová V, Reymond J-L, Engkvist O, Bjerrum E (2020) AiZynthFinder: a fast, robust and flexible open-source software for retrosynthetic planning. J Cheminformatics 12: 70

  11. Thakkar A, Kogej T, Reymond J-L, Engkvist O, Bjerrum EJ (2019) Datasets and their influence on the development of computer assisted synthesis planning tools in the pharmaceutical domain. Chem Sci 11(1):154–168

    Article  Google Scholar 

  12. Segler MHS, Preuss M, Waller MP (2018) Planning chemical syntheses with deep neural networks and symbolic AI. Nature 555(7698):604–610

    Article  CAS  Google Scholar 

  13. Wang X, Qian Y, Gao H, Coley CW, Mo Y, Barzilay R, Jensen KF (2020) Towards efficient discovery of green synthetic pathways with Monte Carlo tree search and reinforcement learning. Chem Sci 11(40):10959–10972

    Article  CAS  Google Scholar 

  14. Lin K, Xu Y, Pei J, Lai L (2020) Automatic retrosynthetic route planning using template-free models. Chem Sci 11(12):3355–3364

    Article  CAS  Google Scholar 

  15. Wang Z, Zhang W, Liu B (2021) Computational analysis of synthetic planning: past and future. Chin J Chem 39(11):3127–3143

    Article  CAS  Google Scholar 

  16. Sutskever I, Vinyals O, Le QV (2014) Sequence to sequence learning with neural networks. Adv Neural Inf Process Syst 27

  17. Bertz SH (1981) The first general index of molecular complexity. J Am Chem Soc 103(12):3599–3601

    Article  CAS  Google Scholar 

  18. Bertz SH (1983) On the complexity of graphs and molecules. Bull Math Biol 45(5):849–855

    Article  CAS  Google Scholar 

  19. Barone R, Chanon M (2001) A new and simple approach to chemical complexity. Application to the synthesis of natural products. J Chem Inf Comput Sci 41(2):269–272

    Article  CAS  Google Scholar 

  20. Boda K, Seidel T, Gasteiger J (2007) Structure and reaction based evaluation of synthetic accessibility. J Comput Aided Mol Des 21(6):311–325

    Article  CAS  Google Scholar 

  21. Ertl P, Schuffenhauer A (2009) Estimation of synthetic accessibility score of drug-like molecules based on molecular complexity and fragment contributions. J Cheminformatics 1:8

  22. Voršilák M, Kolář M, Čmelo I, Svozil D (2020) SYBA: Bayesian estimation of synthetic accessibility of organic compounds. J Cheminformatics 12:35

  23. Yu J, Wang J, Zhao H, Gao J, Kang Y, Cao D, Wang Z, Hou T (2022) Organic compound synthetic accessibility prediction based on the graph attention mechanism. J Chem Inf Model 62(12):2973–2986

    Article  CAS  Google Scholar 

  24. Coley CW, Rogers L, Green WH, Jensen KF (2018) SCScore: synthetic complexity learned from a reaction corpus. J Chem Inf Model 58(2):252–261

    Article  CAS  Google Scholar 

  25. Thakkar A, Chadimová V, Bjerrum EJ, Engkvist O, Reymond J-L (2021) Retrosynthetic accessibility score (RAscore)—rapid machine learned synthesizability classification from AI driven retrosynthetic planning. Chem Sci 12(9):3339–3349

    Article  CAS  Google Scholar 

  26. Li B, Chen H (2022) Prediction of compound synthesis accessibility based on reaction knowledge graph. Molecules 27(3):1039

    Article  CAS  Google Scholar 

  27. Liu C-H, Korablyov M, Jastrzebski S, Włodarczyk-Pruszyński P, Bengio Y, Segler M (2022) Retrognn: fast estimation of synthesizability for virtual screening and de novo design by learning from slow retrosynthesis software. J Chem Inf Model 62(10):2293–2300

    Article  CAS  Google Scholar 

  28. Genheden S, Bjerrum E (2022) PaRoutes: towards a framework for benchmarking retrosynthesis route predictions. Digit Discov 1(4):527–539

    Article  Google Scholar 

  29. Bonnet P (2012) Is chemical synthetic accessibility computationally predictable for drug and lead-like molecules? A comparative assessment between medicinal and computational chemists. Eur J Med Chem 54:679–689

    Article  CAS  Google Scholar 

  30. Baba Y, Isomura T, Kashima H (2018) Wisdom of crowds for synthetic accessibility evaluation. J Mol Graph Model 80:217–223

    Article  CAS  Google Scholar 

  31. Rogers D, Hahn M (2010) Extended-connectivity fingerprints. J Chem Inf Model 50(5):742–754

    Article  CAS  Google Scholar 

  32. Hassan M, Brown RD, Varma-O’Brien S, Rogers D (2006) Cheminformatics analysis and learning in a data pipelining environment. Mol Divers 10(3):283–299

    Article  CAS  Google Scholar 

  33. Kim S, Chen J, Cheng T, Gindulyte A, He J, He S, Li Q, Shoemaker BA, Thiessen PA, Yu B, Zaslavsky L, Zhang J, Bolton EE (2021) PubChem in 2021: new data content and improved web interfaces. Nucleic Acids Res 49(D1):1388–1395

    Article  Google Scholar 

  34. RDKit: Open-source cheminformatics. https://rdkit.org/

  35. Voršilák M, Svozil D (2017) Nonpher: computational method for design of hard-to-synthesize structures. J Cheminformatics 9:20

    Article  Google Scholar 

  36. McCulloch WS, Pitts W (1943) A logical calculus of the ideas immanent in nervous activity. Bull Math Biophys 5(4):115–133

    Article  Google Scholar 

  37. Lawson A.J, Swienty-Busch J, Géoui T, Evans D (2014) Chap. 8. The making of Reaxys—towards unobstructed access to relevant chemistry information. In: the future of the history of chemical information. ACS symposium series, vol 1164, pp 127–148. American Chemical Society, Washington, DC

  38. Morgan HL (1965) The generation of a unique machine description for chemical structures-a technique developed at chemical abstracts service. J Chem Doc 5(2):107–113

    Article  CAS  Google Scholar 

  39. Gaulton A, Hersey A, Nowotka M, Bento AP, Chambers J, Mendez D, Mutowo P, Atkinson F, Bellis LJ, Cibrián-Uhalte E, Davies M, Dedman N, Karlsson A, Magariños MP, Overington JP, Papadatos G, Smit I, Leach AR (2017) The ChEMBL database in 2017. Nucleic Acids Res 45(D1):945–954

    Article  Google Scholar 

  40. Friedman JH (2001) Greedy function approximation: a gradient boosting machine. Ann Stat 29(5):1189–1232

    Article  Google Scholar 

  41. Kocsis L, Szepesvári C (2006) Bandit based Monte-Carlo planning. In: Fürnkranz J, Scheffer T, Spiliopoulou M (eds) Machine learning: ECML 2006. Lecture notes in computer Science. Springer, Berlin, Heidelberg, pp 282–293

    Chapter  Google Scholar 

  42. Coulom R (2007) Efficient selectivity and backup operators in Monte-Carlo tree search. In: van den Herik HJ, Ciancarini P, Donkers HHLMJ, eds. Computers and games. Lecture notes in computer science, pp 72–83. Springer, Berlin, Heidelberg

  43. Chaslot GMJ-B, Winands MHM, Herik HJVD, Uiterwijk JWHM, Bouzy B (2008) Progressive strategies for Monte-Carlo tree search. New Math Natural Comput 04(03):343–357

    Article  Google Scholar 

  44. Weininger D (1988) SMILES, a chemical language and information system. 1. Introduction to methodology and encoding rules. J Chem Inf Model 28(1):31–36

    Article  CAS  Google Scholar 

  45. Skoraczyński G, Dittwald P, Miasojedow B, Szymkuć S, Gajewska EP, Grzybowski BA, Gambin A (2017) Predicting the outcomes of organic reactions via machine learning: are current descriptors sufficient? Sci Rep 7(1):3582

    Article  Google Scholar 

  46. Medina-Franco JL (2021) Grand challenges of computer-aided drug design: the road ahead. Front Drug Discov 1:728551

    Google Scholar 

  47. Jiang D, Wu Z, Hsieh C-Y, Chen G, Liao B, Wang Z, Shen C, Cao D, Wu J, Hou T (2021) Could graph neural networks learn better molecular representation for drug discovery? A comparison study of descriptor-based and graph-based models. J Cheminformatics 13:12

  48. Spearman C (1904) The proof and measurement of association between two things. Am J Psychol 15(1):72–101

    Article  Google Scholar 

  49. Student (1908) The probable error of a mean. Biometrika 6(1): 1–25

Download references

Acknowledgements

Not applicable.

Funding

GS was supported by Polish National Science Center grant number 2019/33/N/ST6/02949. Partially GS and BM were supported by Polish National Science Center grant number 2018/31/B/ST1/00253.

Author information

Authors and Affiliations

Authors

Contributions

GS performed statistical data analysis and revised a tool for analysis of synthetic accessibility scores and AiZynthFinder search trees. MK implemented a tool for analysis of synthetic accessibility scores and AiZynthFinder search trees. BM conceived the idea of the project and proposed an analysis procedure. AG supervised the project and discussed the results. All authors co-wrote the manuscript. All authors read and approved the final manuscript

Corresponding author

Correspondence to Grzegorz Skoraczyński.

Ethics declarations

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Additional file 1.

 A Microsoft Excel spreadsheet containing a database of analyzed molecules with their SMILES encoding and synthesis information.

Additional file 2: Figure S1.

ROC curves for discrimination of solved and not solved nodes bysynthetic accessibility scores. Table S1. Tree max depth for replacing a fraction of the reward with anappropriately scaled synthetic accessibility score (SAscore, SCScore, SYBA). Table S2. Tree maximum width for replacing a fraction of the reward with anappropriately scaled synthetic accessibility score (SAscore, SCScore, SYBA). Table S3. Tree node count for replacing a fraction of the reward with anappropriately scaled synthetic accessibility score (SAscore, SCScore, SYBA). Table S4. Number of not solved leaves for replacing a fraction of the reward withan appropriately scaled synthetic accessibility score (SAscore, SCScore, SYBA).

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Skoraczyński, G., Kitlas, M., Miasojedow, B. et al. Critical assessment of synthetic accessibility scores in computer-assisted synthesis planning. J Cheminform 15, 6 (2023). https://doi.org/10.1186/s13321-023-00678-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13321-023-00678-z

Keywords