Abstract

Drug combinations are required to treat advanced cancers and other complex diseases. Compared with monotherapy, combination treatments can enhance efficacy and reduce toxicity by lowering the doses of single drugs—and there especially synergistic combinations are of interest. Since drug combination screening experiments are costly and time-consuming, reliable machine learning models are needed for prioritizing potential combinations for further studies. Most of the current machine learning models are based on scalar-valued approaches, which predict individual response values or synergy scores for drug combinations. We take a functional output prediction approach, in which full, continuous dose-response combination surfaces are predicted for each drug combination on the cell lines. We investigate the predictive power of the recently proposed comboKR method, which is based on a powerful input–output kernel regression technique and functional modeling of the response surface. In this work, we develop a scaled-up formulation of the comboKR, which also implements improved modeling choices: we (1) incorporate new modeling choices for the output drug combination response surfaces to the comboKR framework, and (2) propose a projected gradient descent method to solve the challenging pre-image problem that is traditionally solved with simple candidate set approaches. We provide thorough experimental analysis of comboKR 2.0 with three real-word datasets within various challenging experimental settings, including cases where drugs or cell lines have not been encountered in the training data. Our comparison with synergy score prediction methods further highlights the relevance of dose-response prediction approaches, instead of relying on simple scoring methods.

Introduction

Drug combination therapies are increasingly needed for treating cancer and other complex diseases [1–3]. A number of studies have used various computational tools of bioinformatics for a better understanding of the mechanisms of diseases and for the treatment of different diseases. Recently published in silico studies showed that bioinformatics plays a vital role in drug development and better understanding of the molecular basis of diseases [4–8]. However, systematic discovery of synergistic drug combinations remains challenging, due to limited capacity and cost of experimental drug screening. Even with high-throughput screening technologies [9–12], exhaustive testing of pairwise combinations leads to increasingly large combinatorial search space of drug combinations; for instance, considering 100 drugs at five concentrations in 10 cell lines would result in more than 1 million laboratory tests. In this regard, machine learning approaches can speed-up and rationalize the search process by pinpointing the most promising search directions [13, 14].

While most of the machine learning research focuses on prediction of drug combination synergy scores or other single-valued prediction problems [14–18], a few recent approaches consider scalar-valued machine learning approaches for predicting directly the drug combination responses at selected concentrations [19–23]. These methods are more applicable in practice and to various combination prediction tasks beyond synergy, as they are not dependent on any specific synergy metric. Indeed, the different synergy models that score the divergence between the expected and measured combination response—such as the highest single agent (HSA) model [24], Bliss independence model [25], or Loewe additivity model [26]—do not always agree in terms of synergy definition, partly due to large differences in drug concentrations and maximum response values across experimental studies [27]. The approaches that predict individual response values have severe limitations. For instance, predictions for a given drug combination might not be consistent, nor is there any guarantee that they would follow a smooth response surface. Indeed, individual response values can be seen as samples of an underlying target function: a non-linear parametric surface describing the drug interactions. By directly predicting such functional form, the models provide better interpolation and extrapolation along the surfaces, and more consistently capture and quantify potential synergistic patterns.

To this end, [28] and [29] proposed models that predict full drug interaction surfaces, instead of discrete values sampled from them. Of these structured prediction approaches, PIICM introduced in [28] is based on Gaussian process regression, working similarly to a matrix completion approach in filling-in missing drug surface entries from the batch of experiments. ComboKR [29] belongs to the family of functional output regression methods that predict the drug interaction surfaces from primary features, such as gene expression of cell lines or structural fingerprints of drugs.

The comboKR approach is based on a powerful input–output kernel regression (IOKR) framework [30–32], with a novel normalization scheme for the drug interaction surfaces [29]. In this work, we propose an improved formulation for the comboKR framework that both scales up to larger datasets, and obtains a predictive performance superior to the original approach; we call this scalable variant comboKR 2.0. In addition to the improved scalability, we propose two additional improvements to the model: 1) we introduce the usage of difference to the neutral interaction surfaces as the intermediate target for the learner, and 2) we solve the arising pre-image problem, which is often solved with simple candidate set optimization, with a more advanced projected gradient descent algorithm. We perform comprehensive computational experiments to validate and benchmark our approach, showing results in three different drug combination datasets and three predictive scenarios of varying difficulties. We also perform ablation studies, showcasing the relevance of our proposed model improvements, as well as perform comparison against two popular deep learning -based synergy score prediction approaches.

Methods

ComboKR

ComboKR, first introduced in [29], uses generalized kernel dependency estimation [32] or IOKR [30, 31] framework to predict full, continuous drug interaction response surfaces |$y\in \mathcal{Y}$| from the input data |$x\in \mathcal{X}$|⁠. The idea of the framework, in a nutshell, is to cast the difficult difficult structured learning problem |$f:\mathcal{X}\rightarrow \mathcal{Y}$| into a simpler learning problem. This is done by mapping the structured data in |$\mathcal{Y}$| to a reproducing kernel Hilbert space (RKHS) |$\mathcal{H}_{\mathcal{Y}}$| associated with a kernel chosen for the output data, |$k_{y}:\mathcal{Y}\times \mathcal{Y}\rightarrow \mathbf{R}$|⁠. This gives rise to a vector-valued (or function-valued for kernels associated with infinite-dimensional RKHS) learning problem to learn |$g:\mathcal{X}\rightarrow \mathcal{H}_{\mathcal{Y}}$|⁠. After learning |$g$|⁠, a (usually ill-posed) pre-image problem has to be solved to map the predictions back to |$\mathcal{Y}$|⁠.

ComboKR uses kernel ridge regression (KRR) in learning the function |$g$|⁠, with which a closed-form solution is obtained and directly inserted to the final pre-image problem. With this, and assuming a normalized output kernel |$k_{y}$| (i.e. |$k_{y}(y,y)=1 \; \forall y\in \mathcal{Y}$|⁠), the final optimization problem to obtain predictions can be written as

(1)

in which |$\mathbf{K}_{x}$| is the |$n\times n$| kernel matrix for |$n$| input training samples; |$\mathbf{I}_{n}$| is identity matrix of same size; |$\lambda $| is the KRR regularisation parameter; and the shorthand |$k_{y}(y,Y)$| refers to the vector |$[k_{y}(y, y_{1}),..., k_{y}(y, y_{n})]$| with |$\{y_{i}\}_{i=1}^{n}=Y$| the outputs of the training set; |$k_{x}(X,x)$| is defined analogously. It is good to note that here the kernel trick is used for both input and output data. For the output data, a crucial part of the comboKR approach was to propose a novel drug interaction surface normalization scheme within the output kernel |$k_{y}$|⁠, which captures the relevant parts for the similarity comparisons with it. It was demonstrated that this approach was much more effective than the use of a more traditional kernel computation.

The search space |$\mathcal{Y}$| in pre-image problem is vast, and the problem is ill-posed [33, 34]. ComboKR solved the problem with candidate set optimization: instead of considering the full |$\mathcal{Y}$|⁠, the search is restricted to finding maximum value over a smaller candidate set |$C_{y}$|⁠. Every element in the set is tried out, and the one giving maximum value to 1 is chosen as the final prediction.

As is usual for kernel methods, the solution depends on inverting a |$n\times n$| matrix, |$\mathbf{K}_{x}+\mathbf{I}$|⁠, with |$n$| the number of training samples. In general the input data space would consist of triplets |$(x_{c}, x_{d1}, x_{d2})\in \mathcal{X}=\mathcal{X}_{c}\times \mathcal{X}_{d}\times \mathcal{X}_{d}$|⁠, with cell line features and the features from the two drugs in the combination. In comboKR, the model was built for one cell line at a time, and input data consisted of tuples |$(x_{d1}, x_{d2})\in \mathcal{X}_{d}\times \mathcal{X}_{d}$|⁠. Now |$\mathbf{K}_{x}=\mathbf{R}[\mathbf{K}_{d}\otimes \mathbf{K}_{d}]\mathbf{R}^\top $|⁠, with |$\mathbf{K}_{d}\in \mathbf{R}^{n_{d}\times n_{d}}$| begin kernel matrix of |$n_{d}$| drug features, and |$\mathbf{R}\in \{0, 1\}^{n\times (n_{d}\times n_{d})}$| selects the |$n$| training samples from the full Kronecker product matrix. For the full problem including also cell lines, |$\mathbf{K}_{x}=\mathbf{R^{\prime}}[\mathbf{K}_{c}\otimes \mathbf{K}_{d}\otimes \mathbf{K}_{d}]\mathbf{R^{\prime}}^\top $| with |$\mathbf{R^{\prime}}\in \{0, 1\}^{n\times (n_{c} \times n_{d}\times n_{d})}$|⁠, which is not practical to compute for most datasets.

Scaling up ComboKR

At the first glance, scaling up the inverse computation might seem trivial—in the end the kernel matrix on the full data is based on simple Kronecker product |$\mathbf{K}_{c}\otimes \mathbf{K}_{d}\otimes \mathbf{K}_{d}$|⁠. However the training set is rarely complete, and |$\mathbf{K}_{x}$| does not include a full set of cell–drug–drug combinations. In practice, for training one gets a sampled version of this matrix, which does not enjoy the efficient inverse formula taking advantage of the Kronecker product structure (eq. (512) in [35]).

A solution is to use efficient iterative solvers for the KRR problem. Such solver has been proposed in [36, 37], with practical implementation in [38]. This iterative generalized vec-trick solver is able to take advantage of the sampled Kronecker product structure of the input kernel matrix, and solve a linear equation of the form

by efficient computation of the iterative update |$\mathbf{a} \leftarrow \mathbf{R}(\mathbf{A}\otimes \mathbf{B})\mathbf{R}^\top \mathbf{c}$| with sparse |$\mathbf{R}$|⁠. However, the method is only proposed for products consisting of two matrices in the product. Thus with the drug response prediction application, a part of the three-part Kronecker product has to be explicitly computed before applying the solver; |$\mathbf{K}_{c}\otimes \mathbf{K}_{d}$| or |$\mathbf{K}_{d}\otimes \mathbf{K}_{d}$|⁠, depending on if number of cell lines or drugs in the data is smaller.

Using the iterative solver to obtain the KRR solution for function |$g:\mathcal{X}\rightarrow \mathcal{H}_{\mathcal{Y}}$| means that it is no longer possible to take advantage of the kernel trick in the output space: now for a data sample |$x$| the prediction step is done by solving the pre-image problem

(2)

where |$\psi $| denotes the feature map associated with the normalized output kernel |$k_{y}$|⁠. For some kernels, such as the simple linear kernel, |$\psi (y)$| is easily computable, but e.g. for the RBF kernel used in comboKR, the feature space is infinite-dimensional. In order to be able to compute (2), we need to consider approximation of this output space, |$\tilde{\mathcal{H}}_{\mathcal{Y}}$|⁠, and learn with KRR instead |$\tilde{g}:\mathcal{X}\rightarrow \tilde{\mathcal{H}}_{\mathcal{Y}}$|

To this end, we consider a Nyström approximated feature map of the output surface kernel |$k_{y}$|⁠. Generating a diverse set of landmark surfaces, |$L_{y} = \{l_{i}\}_{i=1}^{n_{L}}$|⁠, which cover the space well and give a good quality approximation, we can compute |$\tilde{\psi }(y) = k_{y}(y, L_{y})k_{y}(L_{y}, L_{y})^{-1/2}$|⁠, giving the final pre-image problem as

(3)

Most of the drug combinations in the datasets exhibit behaviour close to additive effects, i.e. not antagonistic or synergistic combinations. We can write a generic surface function as |$S(c, d_{1}, d_{2}, \kappa ) \rightarrow \mathbb{R}$|⁠, depending on the cell line and the two drugs (implicitly encoding the monotherapy behaviour) and some interaction parametrization |$\kappa $|⁠. Intuitively, when looking at the resulting surfaces, any small changes in the interaction pattern can be drowned into the more prominent overall patterns of the surfaces. Since the additive part can be directly deduced from the monotherapy responses, there is no need to use the capacity of the learner to learn it, but it can rather be focused on only the interaction pattern. Thus, in order to be able to better capture the tails of the uneven target distribution (i.e. the antagonistic or synergistic surfaces), we propose to write |$S(c, d_{1}, d_{2}, \kappa ) = S(c, d_{1}, d_{2}, 0) + \Delta $|⁠, and focus predicting only |$\Delta $|⁠, i.e. the difference of the true and the neutral interaction surfaces. The final prediction is obtained by adding the neutral surface |$S_{0}(c, d_{1}, d_{2}, 0)$| back to the predicted |$\Delta $|⁠. This is similar modeling choice as in [28], where the surfaces were also partitioned to a non-interaction term and an interaction term that captured the possible synergistic or antagonistic effect.

When solving the pre-image problem, the simplest approach is again to use the candidate set, and try out different interaction parametrizations |$\kappa $| in |$S(c, d_{1}, d_{2}, \kappa )$|⁠. In addition to this, we consider projected gradient descent algorithm. More concretely, we use Nesterov accelerated Adam optimization strategy, and project the result back to valid surfaces by fitting the surface model to the end-result of the gradient descent. Due to the rarity of antagonistic or synergistic interactions, a good first guess for the prediction is a surface without any interaction effects. Thus, the gradient descent uses either |$S(c, d_{1}, d_{2}, 0)$| or zero-valued function as the initial starting point, depending if the learning is done directly on full surfaces, or only on interaction part (⁠|$\Delta $|⁠) of it. The algorithms are displayed in Algorithm 1.

The comboKR 2.0 pipeline is illustrated in Fig. 1, where also the proposed improvements compared with the original comboKR approach are highlighted.

Illustration of the comboKR 2.0 approach to solve the function-valued drug surface prediction problem: (1) first the surface model is fitted to the training data, and landmarks are generated in order to form the required approximation of the output space; (2) the comboKR model is trained by using the generalized vec trick solver; (3) predicting requires solving the challenging pre-image problem; the improvements compared with the previous comboKR approach are highlighted with teal-coloured text.
Figure 1

Illustration of the comboKR 2.0 approach to solve the function-valued drug surface prediction problem: (1) first the surface model is fitted to the training data, and landmarks are generated in order to form the required approximation of the output space; (2) the comboKR model is trained by using the generalized vec trick solver; (3) predicting requires solving the challenging pre-image problem; the improvements compared with the previous comboKR approach are highlighted with teal-coloured text.

Datasets

We consider three datasets–Jaaks et al [11], NCI-ALMANAC [10] and O’Neil et al [39]—collecting drug combination dose-response measurements on various cancer cell lines. The details of the various datasets are described below, while a summary of relevant statistics of them is displayed in Tables 1 and 2.

Table 1

The number of cell lines and drugs in the datasets considered in this work; the table reports the amount and sizes of the drug combination matrices measured in the data—the percentage of surfaces is calculated over all possible combinations (in Jaaks data a surface count does not differentiate between anchor and library drugs)

Dataset#Cell lines#Drugs#SurfacesMeasurements
Jaaks et al12564 68 982 (⁠|$\approx 27.4\%$|⁠)7|$\times $|2; one or two
NCI-ALMANAC60104311 527 (⁠|$\approx 96.9\%$|⁠)3|$\times $|3; one set
O’Neil et al3938 22 527 (⁠|$\approx 82.2\%$|⁠)4|$\times $|4; four sets
Dataset#Cell lines#Drugs#SurfacesMeasurements
Jaaks et al12564 68 982 (⁠|$\approx 27.4\%$|⁠)7|$\times $|2; one or two
NCI-ALMANAC60104311 527 (⁠|$\approx 96.9\%$|⁠)3|$\times $|3; one set
O’Neil et al3938 22 527 (⁠|$\approx 82.2\%$|⁠)4|$\times $|4; four sets
Table 1

The number of cell lines and drugs in the datasets considered in this work; the table reports the amount and sizes of the drug combination matrices measured in the data—the percentage of surfaces is calculated over all possible combinations (in Jaaks data a surface count does not differentiate between anchor and library drugs)

Dataset#Cell lines#Drugs#SurfacesMeasurements
Jaaks et al12564 68 982 (⁠|$\approx 27.4\%$|⁠)7|$\times $|2; one or two
NCI-ALMANAC60104311 527 (⁠|$\approx 96.9\%$|⁠)3|$\times $|3; one set
O’Neil et al3938 22 527 (⁠|$\approx 82.2\%$|⁠)4|$\times $|4; four sets
Dataset#Cell lines#Drugs#SurfacesMeasurements
Jaaks et al12564 68 982 (⁠|$\approx 27.4\%$|⁠)7|$\times $|2; one or two
NCI-ALMANAC60104311 527 (⁠|$\approx 96.9\%$|⁠)3|$\times $|3; one set
O’Neil et al3938 22 527 (⁠|$\approx 82.2\%$|⁠)4|$\times $|4; four sets
Table 2

The primary features used with the various datasets, with number of cell lines and drugs for which they are available

DatasetCell line features#AvailableDrug features#Available
Jaaks et alCopy number124/125MACCS62/64
NCI-ALMANACGene expression [20]60/60MACCS104/104
O’Neil et alGene expression Copy number37/39MACCS38/38
DatasetCell line features#AvailableDrug features#Available
Jaaks et alCopy number124/125MACCS62/64
NCI-ALMANACGene expression [20]60/60MACCS104/104
O’Neil et alGene expression Copy number37/39MACCS38/38
Table 2

The primary features used with the various datasets, with number of cell lines and drugs for which they are available

DatasetCell line features#AvailableDrug features#Available
Jaaks et alCopy number124/125MACCS62/64
NCI-ALMANACGene expression [20]60/60MACCS104/104
O’Neil et alGene expression Copy number37/39MACCS38/38
DatasetCell line features#AvailableDrug features#Available
Jaaks et alCopy number124/125MACCS62/64
NCI-ALMANACGene expression [20]60/60MACCS104/104
O’Neil et alGene expression Copy number37/39MACCS38/38

Jaaks et al.

The dataset consists of testing 64 drugs on 125 cell lines, with the cell lines being from breast, colon, or pancreatic tumors. More specifically, for the three tumor types, 56, 26, and 25 unique drugs were considered, respectively. The response data consists of |$7\times 2$| drug combination measurements done with an anchored approach: the anchor drug with pre-selected low and high doses was tested against a range of seven increasing concentrations of the second compound. The matrices with the same anchor and library drugs were considered as additional monotherapy data, and were not included in the surfaces.

For primary features, we collected the corresponding 125 DepMap ids for the cell lines, and downloaded copy number data for 124 of the cell lines. Similarly for drugs, we searched for PubChem ids, with which we collected the SMILES representation for 62 out of 64 drugs, and computed MACCS structural features.

NCI-ALMANAC

The NCI-ALMANAC dataset is based on testing 104 drugs in 60 cell lines from 9 tissue origins. The data consists of |$3\times 3$| drug combination measurements on the cell lines; in total there are 311 527 drug-drug-cell combinations. In addition to the |$3\times 3$| combination responses, the data also collects monotherapy responses for the corresponding concentrations, providing hence |$4\times 4$| response matrices when they are included.

As primary features for the cell lines, we considered the gene expression data from [20]. As primary features for the drugs, we use the MACCS fingerprints calculated from the isometric SMILES representation.

O’Neil

The dataset is built on testing 38 drugs on 39 cell lines representing multiple cancer types. The data consists of |$4\times 4$| drug combination measurements on the cell lines; in total there are 22 527 drug-drug-cell combinations. (In total there are 22 727 combinations in the data, however 200 of them are not measured at a |$4\times 4$| grid, so we excluded those in favour of easy evaluation on regular grid size.) Median response values over the replicate measurements are considered as the ground truth responses for the experiments.

As the primary features for the drugs, the MACCS fingerprints were calculated based on the isometric SMILES representation.

As primary features for the cell lines, we consider the gene expression profile data and copy number from DepMap [40]. These data were not obtained for two of the cell lines.

Surface model fitted to the data

As in [29], in this work, we consider the BRAID (Bivariate Response to Additive Interacting Doses) drug interaction model [41] that builds on the Hill equation [42, 43] and is motivated by the Loewe additivity principle. Indeed, most of the parameters of the model are the Hill equation parameters for the monotherapy behaviour of the two drugs, with just two interaction parameters.

As comboKR heavily depends on good estimates of the underlying drug response surfaces, Fig. 2 displays the optimal fit of the BRAID surfaces achieved to all the available data. The responses range from 0 to either 1 or 100, representing the percent growth of the cells. For better comparison across the different data sources, for this figure the O’Neil dataset responses have been transformed to [0, 100] from the original [0, 1]. With Jaaks data, if a drug combination occurs in two ways in a cell line, only one of them has been included. For O’Neil data, the median of replicate measurements are used as groundtruth.

Density plots of the fitted BRAID functions sampled at the dose-concentrations of the measurements, compared with the measured responses; for O’Neil data, the response values have been transformed from the [0, 1] range to [0, 100] range for better comparison between the datasets, and the RMSE value in parenthesis corresponds to the original smaller response value range [0, 1].
Figure 2

Density plots of the fitted BRAID functions sampled at the dose-concentrations of the measurements, compared with the measured responses; for O’Neil data, the response values have been transformed from the [0, 1] range to [0, 100] range for better comparison between the datasets, and the RMSE value in parenthesis corresponds to the original smaller response value range [0, 1].

Of the datasets considered, O’Neil achieves the best fit in all metrics, and overall has most even distribution of the response values over the full range. In contrast, both NCI-ALMANAC and Jaaks data contain more responses where the drugs do not inhibit the cell growth.

Computational experiments

Predictive scenarios

We consider three predictive scenarios, here listed in order of increasing difficulty:

  • New combo: a prediction is done on a new triplet |$(c, d_{1}, d_{2})$|⁠, where the cell line and both drugs have been seen in the training set as parts of other triplets, but the triplet itself not.

  • New drug: the triplet |$(c, d_{1}, d_{2})$| contains either a novel pair |$(c, d_{1})$| or |$(c, d_{2})$|⁠: one of the drugs in the triplet has not been seen in the training set associated with that cell line.

  • New cell line: at the prediction time, the triplet |$(c, d_{1}, d_{2})$| queried contains a cell line not seen in the training set. However, the drug combinations in test set can appear in the training data.

Competing methods

We compare our proposed scaled-up comboKR with the one originally proposed in [29], and denote the proposed scalable variant as comboKR 2.0 (Code available at https://github.com/aalto-ics-kepaco/comboKR2.0).

Modeling and the drug combination responses as a structured prediction problem is still an emerging topic. Apart from the previous comboKR approach [29], to our knowledge only PIICM [28] has been proposed in this context. While giving promising results in a simple predictive scenario, it is however not suitable in more challenging ones, as it assumes at the test stage that all cell lines and drugs are present in training data. In addition, due to suboptimal scalability to larger datasets, it is not possible to offer meaningful comparisons in the present experiments.

LTR is a polynomial regression method based on latent tensor reconstruction [44] that has been successfully applied to drug combination response prediction [23] as “comboLTR”. In its original form, it is applicable to settings with one data view, or with multiple views they can be concatenated (early fusion approach). However, this ‘one-view’ approach is not scalable to large datasets. Thus, a more scalable variant has been proposed in [45], which allows to handle magnitudes larger data sets than the original version. We note that even with same parameters, the polynomials learned by the one-view and multi-view comboLTR models are not the same, as the interactions are modeled differently. Due to the scalability issues of the one-view LTR, we use it here only with the O’Neil dataset. Multi-view LTR is considered with all the datasets.

As a simple baseline approach, we compare with BRAID surfaces without any synergy or antagonism. These surfaces are assigned without any learning, and they are based on the same monotherapy Hill equation parameters that the comboKR/comboKR 2.0 uses.

While our setting is designed for predicting response surfaces, we additionally compare our method with two synergy score prediction methods: DeepSynergy [17] and MatchMaker [18]. Both of these models are popular baseline methods for the drug combination synergy prediction tasks based on feedforward neural network architectures; the DeepSynergy model was one of the first deep learning approaches proposed in the synergy score prediction literature.

In our comparison with the synergy score prediction models, we predict the response surfaces with comboKR and comboLTR as usual, and subsequently compute the synergy scores as a postprocessing step. DeepSynergy and MatchMaker are directly trained with synergy scores; we consider both Bliss and Loewe synergies separately. In order to have only one synergy score per surface, we take the mean of the synergy scores of the available combinations dose-responses (e.g. for O’Neil dataset over the 16 combo measurements of the surface, similarly for the other data sets; see Table 1).

Experimental details

ComboKR relies on kernels on both inputs and outputs. As mentioned, for outputs we consider RBF kernel on the normalized surfaces sampled on a grid. For the input features, either Tanimoto or RBF kernel is calculated—Tanimoto is used for the fingerprint features, RBF kernel for others. In all RBF kernels, the kernel parameter |$\sigma $| is set as mean of pairwise distances in the data.

Pearson correlation results of the various methods on the three datasets, between the measured and predicted drug combination responses; the horizontal dashed lines highlight the range of baseline results of predicting neutral interaction surface without learning via the BRAID function (last entries in the plots), and on top of the plots the P-value results of Wilcoxon rank-sum test (”WRS”) are shown (with Bonferroni correction), compared with our comboKR 2.0.
Figure 3

Pearson correlation results of the various methods on the three datasets, between the measured and predicted drug combination responses; the horizontal dashed lines highlight the range of baseline results of predicting neutral interaction surface without learning via the BRAID function (last entries in the plots), and on top of the plots the P-value results of Wilcoxon rank-sum test (”WRS”) are shown (with Bonferroni correction), compared with our comboKR 2.0.

For non-kernel based approaches (i.e. ComboLTR), the empirical features based on the input kernels are considered. Additionally, as a scalar-valued prediction approach, ComboLTR requires concentration values. These are given as one-hot encoded vectors.

Cross-validation is performed to select the regularisation parameters for the models. For the various methods, the parameters searched over are as follows:

  • comboKR 2.0: we consider early stopping of the iterative generalized vec-trick solver as the regularisation; the number of iterations are cross-validated over [3, 5, 10]. Number of landmark BRAID surfaces in Nyström approximated output feature map can also been as a regularizer in all the models. It’s decided by the number of eigenvalues greater than or equal to 0.1 in the kernel formed from generated landmark surfaces. This results in only 27 landmark surfaces in Nyström approximation. Unless otherwise stated, we consider difference surface modeling and projected gradient descent as the pre-image solver.

  • Original comboKR: regularisation parameter |$\lambda $| in IOKR: [1e-2, 5e-2, 1e-1, 5e-1]

  • ComboLTR: in both variants order of the polynomial is chosen from [3, 5], and rank from [10, 20].

  • We perform the cross-validation for DeepSynergy and MatchMaker over learning rate, and cross-validate over learning rates [0.01, 0.001, 0.0001, 0.00001].

Results

Overall comparison

We present the overall Pearson correlation results of the compared approaches in Fig. 3 (see Supplementary material for these results in tabular format, as well as results about Spearman correlation, as well as Pearson correlation results on calculated Bliss and Loewe synergy scores). The comboKR 2.0 outperforms the previous comboKR approach; only with O’Neil data and on the easiest new combo scenario, the original comboKR perform better than 2.0. The major improvement from the original is seen especially clearly in the new drug experimental scenario, where the comboKR 2.0 outperforms the baseline, while original version obtains worse results than the baseline. Overall, comboKR outperforms the comboLTR in all predictive scenarios. As can be expected, the new cell line scenario is very challenging, and the difference of comboKR 2.0 to the baseline is very small. The original ComboKR model is trained cell-by-cell manner, while the other methods are trained over multiple cell lines: thus, comboKR is not applicable in the new cell line scenario.

In order to perform statistical testing on the results, we computed for each surface prediction its root mean squared error (RMSE) to the groundtruth surface measurements. With these, we performed Wilcoxon rank-sums test, as well as Kolmogorov-Smirnov test, comparing the other methods with our proposed comboKR 2.0. We display the Bonferroni-corrected P-values as part of Fig. 3. Supplementary material shows additional results for the pairwise comparison of the methods w.r.t RMSE with the groundtruth surface measurements.

Comparison with synergy score prediction methods

As drug combination synergy prediction methods outnumber the drug combination response prediction methods, we here present comparison between the two types of methods on the O’Neil dataset. Here, we consider two types of synergy models: Loewe and Bliss. We present the Pearson correlation results between the predicted synergy score and the synergy score computed from the ground truth data in Fig. 4 (these results are given in tabular format in the supplementary material). There, we also show the results of the Wilcoxon rank-sum test with Bonferroni correction, where, for all the methods, we compared the distributions of the predicted scores with the ground truth scores, taking the median over the five splits. We note that for the dose-response prediction models (comboKR, comboLTR), the synergy score prediction is obtained via a post-processing step, where the scores are calculated based on the predicted responses.

Pearson correlations of the synergy score prediction results on O’Neil data, compared with the scores computed from the groundtruth measurements; on top of the plots them median (over the five splits) P-value results of Bonferroni-corrected Wilcoxon rank-sum test are shown, for when the synergy score predictions are compared with the ground truth scores.
Figure 4

Pearson correlations of the synergy score prediction results on O’Neil data, compared with the scores computed from the groundtruth measurements; on top of the plots them median (over the five splits) P-value results of Bonferroni-corrected Wilcoxon rank-sum test are shown, for when the synergy score predictions are compared with the ground truth scores.

The results highlight the relevance of dose-response prediction methods. While the synergy score prediction method MatchMaker is competitive in the new combo experimental setting with Bliss synergy score, the same method performs much worse when Loewe is considered. Indeed, it seems that Loewe synergy scores are much more difficult than Bliss for these methods. Deep learning methods often thrive with large data sets sizes; it could be expected that the performance of DeepSynergy and MatchMaker would improve with larger data set sizes, or with more detailed cross-validation strategies than our comparison considered.

Cell line feature importance

In addition to the importance of even including the cell line features, we briefly investigate the performance with different cell line features with O’Neil dataset. To this end, we consider using three different primary features for the cell lines: A) the gene expression and copy number features used in other experiments; B) one-hot encoding of the cell line id; and C) one hot encoding of the cell line id, concatenated with one-hot encoding of the cancer type id.

The results are displayed in Fig. 5. In the easier new combo training scenario all methods benefit from using id features; comboLTR seems to additionally benefit from using the cell type id features. For comboKR 2.0, the performance of using cell type and cell line ids is here equal to that using more complicated primary features. The new drug scenario starts to show how the complicated primary features can be a better choice than the id features, while the difference between the features in comboLTR also decreases.

Pearson correlations of the response predictions with various cell line features on O’Neil data: A: primary features (gene expression and copy number features) B: one-hot encoded cell line id C: one-hot encoded cell id with one-hot encoded cancer type id.
Figure 5

Pearson correlations of the response predictions with various cell line features on O’Neil data: A: primary features (gene expression and copy number features) B: one-hot encoded cell line id C: one-hot encoded cell id with one-hot encoded cancer type id.

Ablation study

The scaled-up version of comboKR often outperforms the original implementation that considers each cell line a separate experiment. In order to investigate how much of the gain/loss is due to larger training set sizes, the scaled-up algorithms were run also in cell-by-cell manner and compared with the full runs.

The Pearson correlation of the full and cell-by-cell comboKR 2.0 approach are shown in Table 3. It seems clear that the improved modeling choices improved the comboKR 2.0 approach, independently from scaling up for larger data set sizes. Furthermore, the results indicate that it is possible to obtain good results with comboKR 2.0 even when only data available is from the investigated cell line—an important improvement for practical applications.

Table 3

Pearson correlation results (the mean |$\pm $| standard deviation) on the comboKR 2.0 when trained and tested on full training set, compared with the cell-by-cell setting, and original comboKR

Dataset comboKR 2.0c.KR 2.0 cbcc.KR orig.
JaaksNew combo0.894 |$\pm $| 0.0010.895 |$\pm $| 0.0000.890 |$\pm $| 0.001
New drug0.893 |$\pm $| 0.0040.889 |$\pm $| 0.0040.839 |$\pm $| 0.007
NCI-ALMANACNew combo0.902 |$\pm $| 0.0000.902 |$\pm $| 0.0000.900 |$\pm $| 0.000
New drug0.904 |$\pm $| 0.0030.904 |$\pm $| 0.0030.889 |$\pm $| 0.003
O’NeilNew combo0.939 |$\pm $| 0.0010.940 |$\pm $| 0.0010.938 |$\pm $| 0.001
New drug0.937 |$\pm $| 0.0020.934 |$\pm $| 0.0020.907 |$\pm $| 0.002
Dataset comboKR 2.0c.KR 2.0 cbcc.KR orig.
JaaksNew combo0.894 |$\pm $| 0.0010.895 |$\pm $| 0.0000.890 |$\pm $| 0.001
New drug0.893 |$\pm $| 0.0040.889 |$\pm $| 0.0040.839 |$\pm $| 0.007
NCI-ALMANACNew combo0.902 |$\pm $| 0.0000.902 |$\pm $| 0.0000.900 |$\pm $| 0.000
New drug0.904 |$\pm $| 0.0030.904 |$\pm $| 0.0030.889 |$\pm $| 0.003
O’NeilNew combo0.939 |$\pm $| 0.0010.940 |$\pm $| 0.0010.938 |$\pm $| 0.001
New drug0.937 |$\pm $| 0.0020.934 |$\pm $| 0.0020.907 |$\pm $| 0.002
Table 3

Pearson correlation results (the mean |$\pm $| standard deviation) on the comboKR 2.0 when trained and tested on full training set, compared with the cell-by-cell setting, and original comboKR

Dataset comboKR 2.0c.KR 2.0 cbcc.KR orig.
JaaksNew combo0.894 |$\pm $| 0.0010.895 |$\pm $| 0.0000.890 |$\pm $| 0.001
New drug0.893 |$\pm $| 0.0040.889 |$\pm $| 0.0040.839 |$\pm $| 0.007
NCI-ALMANACNew combo0.902 |$\pm $| 0.0000.902 |$\pm $| 0.0000.900 |$\pm $| 0.000
New drug0.904 |$\pm $| 0.0030.904 |$\pm $| 0.0030.889 |$\pm $| 0.003
O’NeilNew combo0.939 |$\pm $| 0.0010.940 |$\pm $| 0.0010.938 |$\pm $| 0.001
New drug0.937 |$\pm $| 0.0020.934 |$\pm $| 0.0020.907 |$\pm $| 0.002
Dataset comboKR 2.0c.KR 2.0 cbcc.KR orig.
JaaksNew combo0.894 |$\pm $| 0.0010.895 |$\pm $| 0.0000.890 |$\pm $| 0.001
New drug0.893 |$\pm $| 0.0040.889 |$\pm $| 0.0040.839 |$\pm $| 0.007
NCI-ALMANACNew combo0.902 |$\pm $| 0.0000.902 |$\pm $| 0.0000.900 |$\pm $| 0.000
New drug0.904 |$\pm $| 0.0030.904 |$\pm $| 0.0030.889 |$\pm $| 0.003
O’NeilNew combo0.939 |$\pm $| 0.0010.940 |$\pm $| 0.0010.938 |$\pm $| 0.001
New drug0.937 |$\pm $| 0.0020.934 |$\pm $| 0.0020.907 |$\pm $| 0.002

We further investigated substituting the more advanced projected gradient descent solver on difference surface modeling with the simpler modeling possibilities and algorithms to solve the pre-image problem. The different parametrizations with or without difference surface target, and with the different pre-image approaches. We denote these as follows:

The results are displayed in Fig. 6. It is clear that considering difference surfaces within the comboKR model provides consistently better results than using the original surface as the target. Interestingly, using projected gradient descent as the pre-image solver improves the performance further with difference surfaces, but significantly worsens the results with original surface modeling.

Pearson correlation results of the comboKR 2.0 variants on the three datasets, between the measured and predicted drug combination responses.
Figure 6

Pearson correlation results of the comboKR 2.0 variants on the three datasets, between the measured and predicted drug combination responses.

Discussion

In this work, we have implemented multiple improvements to the recently proposed comboKR method [29], a novel way to predict dose responses as full drug combination surfaces. The improvements include scalability to larger datasets with data from multiple cell lines, as well as improvements in the model configuration and approaches for solving the arising pre-image problem. Due to the combinatorial nature of the drug combination response datasets, the scalability of the predictive methods will be very important in the future: as one or two new drugs or cell lines are added to experiments, the amount of response-surfaces grows in combinatorial fashion.

Our work demonstrates the effectiveness of the improved comboKR 2.0 against the original version, as well as against relevant baseline methods, in many relevant experimental scenarios of varying difficulty. The results on three real-world datasets highlight especially the suitability of the variant with difference surface modeling in various experimental designs, coupled with projected gradient descent pre-image optimization. Our method offers competitive performance compared with the competing comboLTR model as well as the original comboKR, and our ablation study highlights the relevancy of the proposed modeling choices. Indeed, with difference surface modeling, the method is able to take advantage of the projected gradient descent pre-image optimization method, unlike in the original direct surface prediction setting, where candidate set optimization far outperforms gradient descent. While comboKR 2.0 excels in the difficult new drug scenario, improved cell line feature modeling might still lead to better performance in this most challenging scenario of predicting responses in previously unseen cell lines.

As a functional output prediction approach—compared with approaches that predict individual drug combination dose-responses—our approach offers predictions that are consistent across the full dose-response surfaces. Moreover, from the predicted surface it is easy to obtain predictions to any dose-concentrations of interest, or compute any type of synergy measure of interest. As illustrated by our experimental comparison with synergy score prediction methods, our comboKR 2.0 is very competitive in synergy score prediction, even if the objective is to predict responses, not the synergy scores.

While the proposed comboKR 2.0 method seems to outperform the other methods, it is important to note that neutral interaction surfaces from Baseline (i.e. additive BRAID surface set without learning) are often extremely competitive: indeed, strong synergism or antagonism is a relatively rare phenomenon [3, 11, 46, 47]. This is especially well seen with the NCI-ALMANAC dataset, which is much more complete than the other two datasets (see Table 1). These large-scale data are therefore expected to contain much more simple additive surfaces, and we found with this dataset most challenging to outperform the additive baseline.

As the case with the initial comboKR, comboKR 2.0 also heavily relies on good surface model to be used in both training and testing time. While the BRAID model has provided promising results, in the future other surface models, such as MuSyC [48] or generalized Hill-type response surfaces [49], could be investigated in our framework to allow for more diverse surface representations.

Key Points
  • Drug combination response prediction has recently been formulated as a functional output predictions problem, where full continuous dose-response surfaces are predicted.

  • This work focuses on recently proposed comboKR model, and offers improvements to it model formulation, as well as computational scalability.

  • The scaled-up comboKR 2.0 method is evaluated with three real-world datasets, and experimental case where novel drugs or cell lines are encountered in the test data.

  • Our experiments highlighted the suitability of dose-response prediction methods in synergy score prediction task, when the scores are obtained directly from the dose-response predictions.

Acknowledgments

The authors acknowledge the computational resources provided by the Aalto Science-IT project.

Funding

This work was supported by Academy of Finland through the grants [334790, 339421, 345802] (MAGITICS, MASF, and AIB, respectively), as well as the Global Programme by Finnish Ministry of Education and Culture. T.A. was supported by grants from the European Union’s Horizon 2020 Research and Innovation Programme (ERA PerMed CLL-CLUE project), Academy of Finland (grants 326238, 340141, 344698, and 345803), Norwegian Health Authority South-East (grants 2020026 and 2023105), the Norwegian Cancer Society (grants 216104 and 273810), the Cancer Society of Finland, and the Sigrid Jusélius Foundation.

Data availability

The original datasets are available for downloading as follows:

The BRAID functions fitted to the data, as well as the Hill equations and the cross-validation splits of the different experiments are available at zenodo: https://zenodo.org/records/14509938; DOI: 10.5281/zenodo.14509938.

References

1

Das
 
P
,
Delost
 
MD
,
Qureshi
 
MH
. et al.  
A survey of the structures of us fda approved combination drugs
.
J Med Chem
 
2019
 
PMID: 30444362
;
62
:
4265
311
.

2

Pomeroy
 
AE
,
Schmidt
 
EV
,
Sorger
 
PK
. et al.  
Drug independence and the curability of cancer by combination chemotherapy
.
Trends in Cancer
 
2022
;
8
:
915
29
.

3

Plana
 
D
,
Palmer
 
AC
,
Sorger
 
PK
.
Independent drug action in combination therapy: Implications for precision oncology
.
Cancer Discov
 
2022
;
12
:
606
24
.

4

Congxue
 
H
,
Mi
 
W
,
Li
 
F
. et al.  
Optimizing drug combination and mechanism analysis based on risk pathway crosstalk in pan cancer
.
Scientific Data
 
2024
;
11
:
74
.

5

Khan
 
S
,
Zakariah
 
M
,
Rolfo
 
C
. et al.  
Prediction of mycoplasma hominis proteins targeting in mitochondria and cytoplasm of host cells and their implication in prostate cancer etiology
.
Oncotarget
 
2016
;
8
:
30830
43
.

6

Khan
 
S
,
Zakariah
 
M
,
Palaniappan
 
S
.
Computational prediction of mycoplasma hominis proteins targeting in nucleus of host cell and their implication in prostate cancer etiology
.
Tumor Biology
 
2016
;
37
:
10805
13
.

7

Khan
 
S
,
Imran
 
A
,
Khan
 
AA
. et al.  
Systems biology approaches for the prediction of possible role of chlamydia pneumoniae proteins in the etiology of lung cancer
.
PloS One
 
2016
;
11
:e0148530.

8

Wang
 
Y
,
Imran
 
A
,
Shami
 
A
. et al.  
Decipher the helicobacter pylori protein targeting in the nucleus of host cell and their implications in gallbladder cancer: An insilico approach
.
J Cancer
 
2021
;
12
:
7214
.

9

Pemovska
 
T
,
Bigenzahn
 
JW
,
Superti-Furga
 
G
.
Recent advances in combinatorial drug screening and synergy scoring
.
Curr Opin Pharmacol
 
2018
;
42
:
102
10
.

10

Holbeck
 
SL
,
Camalier
 
R
,
Crowell
 
JA
. et al.  
The national cancer institute almanac: A comprehensive screening resource for the detection of anticancer drug pairs with enhanced therapeutic activitynci almanac of approved cancer drug combinations
.
Cancer Res
 
2017
;
77
:
3564
76
.

11

Jaaks
 
P
,
Coker
 
EA
,
Vis
 
DJ
. et al.  
Effective drug combinations in breast, colon and pancreatic cancer cells
.
Nature
 
2022
;
603
:
166
73
.

12

Nair
 
NU
,
Greninger
 
P
,
Zhang
 
X
. et al.  
A landscape of response to drug combinations in non-small cell lung cancer
.
Nat Commun
 
2023
;
14
:
3830
.

13

Weinstein
 
ZB
,
Bender
 
A
,
Cokol
 
M
.
Prediction of synergistic drug combinations
.
Current Opinion in Systems Biology
 
2017
;
4
:
24
8
.
Big data acquisition and analysis |$\bullet $| Pharmacology and drug discovery

14

Lianlian
 
W
,
Wen
 
Y
,
Leng
 
D
. et al.  
Machine learning methods, databases and tools for drug combination prediction
.
Brief Bioinform
 
2021
;
23
.

15

Torkamannia
 
A
,
Omidi
 
Y
,
Ferdousi
 
R
.
A review of machine learning approaches for drug synergy prediction in cancer
.
Brief Bioinform
 
2022
;
23
.

16

Liu
 
H
,
Fan
 
Z
,
Lin
 
J
. et al.  
The recent progress of deep-learning-based in silico prediction of drug combination
.
Drug Discov Today
 
2023
;
28
:103625.

17

Preuer
 
K
,
Lewis
 
RPI
,
Hochreiter
 
S
. et al.  
Deepsynergy: Predicting anti-cancer drug synergy with deep learning
.
Bioinformatics
 
2018
;
34
:
1538
46
.

18

Kuru
 
HI
,
Tastan
 
O
,
Ercument
 
A
. et al.  
Matchmaker: A deep learning framework for drug synergy prediction
.
IEEE/ACM Trans Comput Biol Bioinform
 
2021
;
19
:
2334
44
.

19

Ianevski
 
A
,
Giri
 
AK
,
Gautam
 
P
. et al.  
Prediction of drug combination effects with a minimal set of experiments
.
Nature machine intelligence
 
2019
;
1
:
568
77
.

20

Julkunen
 
H
,
Cichonska
 
A
,
Gautam
 
P
. et al.  
Leveraging multi-way interactions for systematic prediction of pre-clinical drug combination effects
.
Nat Commun
 
2020
;
11
:
6136
.

21

Ling
 
A
,
Stephanie
 
R
,
Huang.
 
Computationally predicting clinical drug combination efficacy with cancer cell line screens and independent drug action
.
Nat Commun
 
2020
;
11
.

22

Nafshi
 
R
,
Lezon
 
TR
.
Predicting the effects of drug combinations using probabilistic matrix factorization
.
Frontiers in Bioinformatics
 
2021
;
1
:
708815
.

23

Wang
 
T
,
Szedmak
 
S
,
Wang
 
H
. et al.  
Modeling drug combination effects via latent tensor reconstruction
.
Bioinformatics
 
2021
;
37
:
i93
i101
.

24

Berenbaum
 
MC
.
What is synergy?
 
Pharmacol Rev
 
1989
;
41
:
93
141
.

25

Bliss
 
CI
.
The toxicity of poisons applied jointly 1
.
Annals of applied biology
 
1939
;
26
:
585
615
.

26

Loewe
 
S
.
The problem of synergism and antagonism of combined drugs
.
Arzneimittelforschung
 
1953
;
3
:
285
90
.

27

Vlot
 
AHC
,
Aniceto
 
N
,
Menden
 
MP
. et al.  
Applying synergy metrics to combination screening data: Agreements, disagreements and pitfalls
.
Drug Discov Today
 
2019
;
24
:
2286
98
.

28

Rønneberg
 
L
,
Kirk
 
PDW
,
Zucknick
 
M
.
Dose–response prediction for in-vitro drug combination datasets: A probabilistic approach
.
BMC bioinformatics
 
2023
;
24
:
161
.

29

Huusari
 
R
,
Wang
 
T
,
Szedmak
 
S
. et al.  
Predicting drug combination response surfaces
.
npj Drug Discovery
2025;
2
:2.

30

Brouard
 
C
,
Szafranski
 
M
,
Buc
 
F d’A
.
Input output kernel regression: Supervised and semi-supervised structured output prediction with operator-valued kernels
.
JMLR
 
2016
;
17
:
np
.

31

Brouard
 
C
,
Shen
 
H
,
Dührkop
 
K
. et al.  
Fast metabolite identification with input output kernel regression
.
Bioinformatics
 
2016
;
32
:
i28
36
.

32

Kadri
 
H
,
Ghavamzadeh
 
M
,
Preux
 
P
.
A generalized kernel approach to structured output learning
. In:
ICML
, pp.
471
9
.
PMLR
,
2013
.

33

Scholkopf
 
B
,
Mika
 
S
,
Burges
 
CJC
. et al.  
Input space versus feature space in kernel-based methods
.
IEEE Trans Neural Netw
 
1999
;
10
:
1000
17
.

34

Weston
 
J
,
Chapelle
 
O
,
Vapnik
 
V
. et al.  
Kernel dependency estimation
.
Advances in neural information processing systems
 
2002
;
15
.

35

Petersen
 
KB
,
Pedersen
 
MS
.
The matrix cookbook
.
Technical University of Denmark
 
2008
;
7
:
510
.

36

Airola
 
A
,
Pahikkala
 
T
.
Fast kronecker product kernel methods via generalized vec trick
.
IEEE transactions on neural networks and learning systems
 
2017
;
29
:
3374
87
.

37

Viljanen
 
M
,
Airola
 
A
,
Pahikkala
 
T
.
Generalized vec trick for fast learning of pairwise kernel models
.
Machine Learning
 
2022
;
111
:
543
73
.

38

Pahikkala
 
T
,
Airola
 
A
.
Rlscore: Regularized least-squares learners
.
Journal of Machine Learning Research
 
2016
;
17
:
1
5
.

39

O’Neil
 
J
,
Benita
 
Y
,
Feldman
 
I
. et al.  
An unbiased oncology compound screen to identify novel combination strategies
.
Mol Cancer Ther
 
2016
;
15
:
1155
62
.

40

Tsherniak
 
A
,
Vazquez
 
F
,
Montgomery
 
PG
. et al.  
Defining a cancer dependency map
.
Cell
 
2017
;
170
:
564
576.e16
.

41

Twarog
 
NR
,
Stewart
 
E
,
Hammill
 
CV
. et al.  
Braid: A unifying paradigm for the analysis of combined drug action
.
Sci Rep
 
2016
;
6
:
25523
.

42

Hill
 
AV
,
Hill
 
A
,
Paganini-Hill
 
A
.
The possible effects of the aggregation of the molecules of hemoglobin on its dissociation curves
.
J Physiol
 
1910
;
40
:
iv
vii
.

43

Goutelle
 
S
,
Maurin
 
M
,
Rougier
 
F
. et al.  
The hill equation: A review of its capabilities in pharmacological modelling
.
Fundam Clin Pharmacol
 
2008
;
22
:
633
48
.

44

Szedmak
 
S
,
Cichonska
 
A
,
Julkunen
 
H
. et al.  
A solution for large scale nonlinear regression with high rank and degree at constant memory complexity via latent tensor reconstruction
 
arXiv preprint arXiv:2005.01538
.
2020
.

45

Armah-Sekum
 
RE
,
Szedmak
 
S
,
Rousu
 
J
.
Protein function prediction through multi-view multi-label latent tensor reconstruction
.
BMC bioinformatics
 
2024
;
25
.

46

Yadav
 
B
,
Wennerberg
 
K
,
Aittokallio
 
T
. et al.  
Searching for drug synergy in complex dose–response landscapes using an interaction potency model
.
Comput Struct Biotechnol J
 
2015
;
13
:
504
13
.

47

Kong
 
W
,
Midena
 
G
,
Chen
 
Y
. et al.  
Systematic review of computational methods for drug combination prediction
.
Comput Struct Biotechnol J
 
2022
;
20
:
2807
14
.

48

Meyer
 
CT
,
Wooten
 
DJ
,
Bishal Paudel
 
B
. et al.  
Quantifying drug combination synergy along potency and efficacy axes
.
Cell systems
 
2019
;
8
:
97
108.e16
.

49

Schindler
 
M
.
Modeling synergistic effects by using general hill-type response surfaces describing drug interactions
.
Sci Rep
 
2022
;
12
:
10524
.

This is an Open Access article distributed under the terms of the Creative Commons Attribution License (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted reuse, distribution, and reproduction in any medium, provided the original work is properly cited.

Supplementary data