Skip to main content
Erschienen in: BMC Medicine 1/2022

Open Access 01.12.2022 | Research article

NeRD: a multichannel neural network to predict cellular response of drugs by integrating multidimensional data

verfasst von: Xiaoxiao Cheng, Chong Dai, Yuqi Wen, Xiaoqi Wang, Xiaochen Bo, Song He, Shaoliang Peng

Erschienen in: BMC Medicine | Ausgabe 1/2022

Abstract

Background

Considering the heterogeneity of tumors, it is a key issue in precision medicine to predict the drug response of each individual. The accumulation of various types of drug informatics and multi-omics data facilitates the development of efficient models for drug response prediction. However, the selection of high-quality data sources and the design of suitable methods remain a challenge.

Methods

In this paper, we design NeRD, a multidimensional data integration model based on the PRISM drug response database, to predict the cellular response of drugs. Four feature extractors, including drug structure extractor (DSE), molecular fingerprint extractor (MFE), miRNA expression extractor (mEE), and copy number extractor (CNE), are designed for different types and dimensions of data. A fully connected network is used to fuse all features and make predictions.

Results

Experimental results demonstrate the effective integration of the global and local structural features of drugs, as well as the features of cell lines from different omics data. For all metrics tested on the PRISM database, NeRD surpassed previous approaches. We also verified that NeRD has strong reliability in the prediction results of new samples. Moreover, unlike other algorithms, when the amount of training data was reduced, NeRD maintained stable performance.

Conclusions

NeRD’s feature fusion provides a new idea for drug response prediction, which is of great significance for precise cancer treatment.
Begleitmaterial
Additional file 1: Table S1. RDKit functions and their descriptions. Table S2. Hyperparameters for NeRD. The adjustment of hyperparameters often has an important impact on the specific data set. Table S3. Hyperparameters for DeepCDR, CDRScan, tCNNS, and GraphDRP. These models are all dual-channel or quasi-dual-channel, so the same method is used to adjust the hyperparameters. Table S4. Hyperparameters for RF. The parameters of the RF framework are few, and the parameter selection is generally to adjust the value of N\_estimators, i.e., the number of decision trees. Table S5. Hyperparameters for SVR. Gamma is the coefficient of kernel functions, only valid for `rbf', `poly', and `sigmod'. The parameter Degree only works for `kernel=poly'. C represents the penalty coefficient of the error term. The larger C is, the greater the degree of penalty for wrongly classified samples. Table S6. Hyperparameters for CNN. What we use here is the one-dimensional convolution function provided by pytorch. Table S7. Hyperparameters for MLP. The number of neurons in each layer is also fine-tuned according to the number of hidden layers. Table S8. Hyperparameters for SRMF. SRMF is a method based on matrix factorization, and its hyperparameters mainly include the dimension of the feature space and the regularization parameters. Table S9. Hyperparameters for VAE+MLP. The number of neurons in each layer is also fine-tuned according to the number of hidden layers. Table S10. Number of data instances corresponds to each data partition in the blind test. Table S11. Dataset comparison. Table S12. Blind test dividing data by similarity. Set1-Set5 are test sets with increasing similarity. Set1 has the lowest similarity and Set5 has the highest similarity. The values are the Pearson correlation coefficients. Table S13. Predicted results for the top 1\% of drug-cell lines. We used the trained NERD model to predict drug cell line pairs without IC50 data in the PRISM database, sorted from small to large according to the predicted IC50 value, and then screened the top 1\% of drug-cell line pairs (altogether 2537 pairs across 383 cancer cell lines and 91 drugs). Table S14. The drug target gene list. Based on the list of drugs obtained from the top 1\% of predicted drug-cell line pairs, we found the target genes for these drugs from the PRISM database.
Hinweise

Supplementary Information

The online version contains supplementary material available at https://​doi.​org/​10.​1186/​s12916-022-02549-0.
Xiaoxiao Cheng and Chong Dai contributed equally to this work.

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Abkürzungen
BN
Batch normalization
CCLE
Cancer cell line encyclopedia
CC p
Pearson correlation coefficient
CC s
Spearman correlation coefficient
CNE
Copy number extractor
CNN
Convolutional neural network
DSE
Drug structure extractor
GCN
Graph convolutional network
GDSC
Genomics of drug sensitivity in cancer
GO
Gene ontology
LR
Linear regression
mEE
miRNA expression extractor
MFE
Molecular fingerprint extractor
MLP
Multilayer perceptron
NCI60
National cancer institute 60
PRISM
Profiling relative inhibition simultaneously in mixture
R 2
R squared
RF
Random forest
RMSE
Root mean square error
SMILES
Simplified molecular input line entry specification
SVR
Support vector regression
SVR-L
linear kernel-based SVR

Background

Due to their heterogeneity, tumors from the same tissue origin and pathologic classification exhibit a high degree of genetic and phenotypic variation in individuals [1]. In practice, this translates to differential reactions to treatment. Therefore, to achieve precision medicine, the genetic background and medical history of patients should be considered [2]. Accurate computational prediction of cancer patients’ responses to drug treatment is essential and meaningful to the achievement of precision medication [3]. However, the lack and inaccessibility of data on cancer patients is the limitation for large-scale computational predictions of drug response. In contrast, cell line-based drug response data are abundant and readily available, providing a basis for drug response prediction. Moreover, using the drug response data of cell lines for drug response prediction is the foundation and the most important step in the realization of precision medicine [4]. Furthermore, the effective integration of various types of drug informatics and multi-omics data presents an opportunity to develop drug response prediction models [5, 6].
With the rapid development of biotechnology and the ongoing progress of sequencing technology, a large amount of multi-omics and pharmacological data has been accumulated [7, 8]. In recent years, data from several large-scale drug screening initiatives have been made available, including Genomics of Drug Sensitivity in Cancer (GDSC) [9], Cancer Cell Line Encyclopedia (CCLE) [10], and the US National Cancer Institute 60 human tumor cell line anticancer drug screen (NCI60) [11]. The GDSC database1 is the largest public resource for information on drug sensitivity in cancer cells and molecular markers of drug response. It currently contains nearly 75,000 items of experimental drug sensitivity data, describing the responses of 138 anti-cancer drugs in nearly 700 cancer cell lines [9]. The CCLE database2 is a compilation of gene expressions, chromosomal copy numbers, and massively parallel sequencing data from 947 human cancer cell lines, covering the responses of 24 drugs in 504 cancer cell lines [10]. NCI60 is an in vitro drug discovery tool developed in the late 1980s, which aims to replace the use of transplantable animal tumors in anti-cancer drug screening and test the drug responses of 52,671 drugs in 60 cancer cell lines [11]. They have helped advance the field of precision medicine. However, these studies either test the cellular response of numerous compounds to a limited number of cell lines (e.g., the NCI60 panel), or of a limited number of tumor compounds to numerous cell lines (e.g., the GDSC project). The ideal study should involve a number of drugs (most non-oncologic) screened in a large panel of genomically featured cell lines to capture the molecular diversity of human cancer [12].
To address this problem, Yu et al. reported a biotechnological method called profiling relative inhibition simultaneously in mixtures (PRISM) [13]. Jin et al. applied this method to 500 cell lines covering 21 types of solid tumors and mapped the first generation of human cancer cell metastases, which validated the reliability of the method [14]. Corsello et al. used this method to build a PRISM drug repurposing resource3 database, for which 4518 drugs were tested for growth-inhibitory activity in 578 human cancer cell lines, i.e., a large-scale drug screening process. They thought that this database could be used to build a drug response prediction model in cancer cell lines, thereby suggesting potentially relevant patient groups [12].
Drug response prediction is a core issue of precision medicine. Benefiting from these public datasets, researchers have developed a variety of effective computational methods to predict drug responses in cancer cell lines, thereby promoting the advancement of anti-cancer drug discovery. The rapid development of machine learning also has had a profound impact on biological and medical applications. Menden et al. first develop cancer pharmaco-omics model using multilayer perceptron (MLP). Menden et al. [15] Ridge regression [16], Lasso regression [17], random forest (RF) [18], and some Bayes-based methods [19, 20] are used to build drug response prediction models. Due to their powerful capabilities in model integration, such algorithms have been used to conduct systematic research on drug response prediction, combined with integrated strategies and multicore multitask learning techniques [2123]. Nevertheless, because of the complexity of multi-omics data, these methods often face the problem of “small n, big p,” i.e., a feature dimension much greater than the number of samples [24]. This makes it difficult for such methods to effectively extract features from complex omics data. Some researchers [2527] focus on feature selection, which is a major antidote to the statistical and computational problems that the high-dimensional omics input data typically entail [28], to improve prediction accuracy on classic machine learning models. Auto-HMM-LMF [29] and Dr.VAE [30] used autoencoders to solve the problem of high dimensionality of omics data, but they did not explore the characteristics of drugs. Effective fusion of multi-omics data is also one of the core issues of drug response prediction. Existing fusion categories can be summarized as early-fusion [15, 31], late-fusion [32] and intermediate-fusion [24, 33, 34]. Intermediate-fusion shows better performance in this problem. In tCNNS [33], a set of twin convolutional neural networks (CNNs) was used to combine the simplified molecular input line entry specification (SMILES) of a drug with the genome mutation data of a cell line. However, the limitations of CNNs render it unable to deal with features of different data structures and dimensions. GraphDRP [24] and DeepCDR [34] extract the drug structure information represented by a graph through a graph convolutional network (GCN). Although they made some progress in model performance, they only used a single drug feature. Furthermore, using multisource information fusion with insufficient data to train the model, to maintain good prediction accuracy poses a challenge. The scarcity of data due to the high cost of labeling remains the main problem in biomedical applications.
In response to the above problems, we propose a multichannel Neural network model to predict the cellular Response of Drugs (NeRD), using the PRISM drug response database. NeRD combines a one-dimensional CNN, stacked autoencoder, and GCN to effectively extract and integrate the global and local structure of a drug, as well as the cell line characteristics from multi-omics data. The fully connected network is then used to predict the final drug response score. Experimental results show that our method can effectively integrate multisource information and combine the features of different data structures and dimensions. NeRD outperformed seven comparison methods on all evaluation metrics on the PRISM database. Moreover, when the amount of training data was reduced, NeRD maintained stable performance and was more robust than the comparison algorithms. We summarized our contributions as follows.
  • An accurate drug response prediction model NeRD is proposed. The model with a multichannel structure can effectively extract the features with different data structures and dimensions and integrate multisource information of drugs and cell lines.
  • The fusion of multisource information makes the model more robust. Unlike other algorithms, when the amount of training data is reduced, NeRD maintains stable performance.
  • We use a recently proposed database PRISM and prove its practicability. The database contains more drug-cell line pairs and is worthy of attention by researchers.

Methods

Database and data preprocessing

The data we use comes from the PRISM drug repurposing database, which contains the IC50 values, i.e., the concentration of a drug required to inhibit 50% of the cell line activity, for 1448 drugs across 480 cell lines. The lower the value the better the drug’s effect. We retrieved the SMILES feature characterizing the overall structure information of all drugs and the molecular fingerprint feature of local structure information. For cell lines, we selected the DNA copy number and miRNA expression data from multiple omics features. A total of 388 cell lines had data on both of the above omics features. The meanings of features and the reasons for selecting them are as follows.

Simplified molecular input line entry specification (SMILES)

An ASCII string represents the three-dimensional chemical structures of drugs. We used the RDKit toolkit [35] to transform a SMILES string to a molecular graph that reflects interactions between atoms inside drugs. Each atom was represented by a node, and the bonds between atoms were represented by edges. And each node contains five types of atom features: atom symbol, atom degree calculated by the number of bonded neighbors and hydrogen atoms, total number of hydrogen atoms, implicit value of the atom, and whether the atom is aromatic. These atom features are encoded into a 78-dimensional binary vector [24]. The RDKit functions we used and their descriptions can be found in Additional file 1: Table S1.

Molecular fingerprint

For 1448 drugs, we extracted chemical structure data in SDF format from the PubChem compound database [36]. Each drug was encoded into an 881-dimensional substructure vector defined in PubChem using the R package ChemmineR. Each drug is represented by a binary fingerprint that indicates the existence of a predefined chemical structure fragment. If a drug contains the corresponding chemical fingerprint, the element is 1, and otherwise it is 0.
The above two drug features were selected to extract the global and local structural features of drugs together, so as to improve the reliability of the results. In this paper, the local structure features of drugs refer to the substructure information of drug molecules represented by molecular fingerprints, because molecular fingerprints describe whether drug molecules contain certain substructures. Then the other drug feature, the molecular map, contains the information of the entire molecular structure, that is, the global structural feature of the drug.

Omics data

We acquired DNA copy number and miRNA expression data from the CCLE database for 338 cell lines. The DNA copy number data consist of 23,316-dimensional vectors that represent the number of occurrences of a specific DNA sequence in a haploid genome, which can reflect the characteristics of cell lines at the gene level. Studies have shown that copy number alterations are ubiquitous in cancer, and many of which are disadvantageous [37]. They are involved in the formation and progression of cancer and contribute to cancer proneness [38]. Analysis of copy number alteration data can help in cancer diagnosis and treatment by providing a better understanding of the biological and phenotypic effects of cancer [39]. Based on these studies, we have also considered this data as feature data for cancer cell lines. The miRNA expression data consist of 734-dimensional vectors. It is a kind of noncoding RNA molecule that can inhibit or degrade mRNA translation by binding to complementary target mRNA. It plays an important role in cell differentiation, proliferation, and survival [40]. Functional studies have confirmed that a causal relationship exists between abnormal miRNA regulations in many cancer cases. miRNAs, as tumor suppressors or oncogenes (oncomiRs), miRNA mimics, and molecules targeting miRNAs (antimiRs), have shown prospects in preclinical development [41].

Data preprocessing

To avoid the adverse effects of the different distributions of the DNA copy number and miRNA expression data on model training, we normalized them before inputting the feature extraction channels. For the drug SMILES (graph) and molecular fingerprint (binary vector), due to the particularity of the data format, we performed no processing before input. The range of values for IC50 is too large, and there are outliers. Therefore, we logarithmically processed the raw data while ensuring that the original IC50 values could be recovered. We also used a box-plot to remove outliers [42]. We took the upper quartile \(Q_{3}\) and lower quartile \(Q_{1}\) of all response data. Then, we got the interquartile range \(IQR=Q_{3}-Q_{1}\). Finally, IC50 values less than \(Q_{1}-1.5\times IQR\) and greater than \(Q_{3}+1.5\times IQR\) were regarded as outliers. Specifically, the data we use contained 1448 drugs and 388 cell lines. Among them, there are 249,784 data with labels (44.46%). After removing the 15,976 outliers counted by the box-plot, we ended up using a total of 233,808 labels.

Multichannel-based neural network

Overview

Due to the different data structures and dimensions of drug features and cell line features, we designed different feature extraction networks for the four types of features (Fig. 1).
We use the SMILES sequence containing global structure information and the molecular fingerprint containing local structure information as drug features. The SMILES sequence describes the three-dimensional chemical structure of drugs. To extract the maximum structural information, we use SMILES in the graph form as the input of the drug structure extractor (DSE). To extract feature information from the graph, we use a method that can perform deep learning on graph data, the GCN, through which we can obtain the structural features of drug molecules. Since the data structure of a graph is different from other features and cannot be directly integrated, global maximum pooling is used to convert the feature data from a matrix to a vector, and its features are normalized to 128 dimensions through a fully connected network. Molecular fingerprints describe whether a drug has certain substructures, and can represent its local structural features. Since the data structure of the molecular fingerprint is a standardized binary vector, it can be directly used as the input of the molecular fingerprint extractor (MFE). Then, we use a one-dimensional CNN to extract the features of these substructures, and normalize them to vectors of the same dimension. The two feature vectors representing a drug are spliced to obtain its final feature representation.
We use miRNA expression data and the DNA copy number as features for cell lines. We designed a miRNA expression extractor (mEE) based on a one-dimensional CNN. We input the feature vector describing the miRNA into this channel and extracted its potential features. However, the DNA copy number cannot be directly extracted by the above neural network model due to its ultra-high dimensionality. So we designed a copy number extractor (CNE) based on a stacked autoencoder and performed nonlinear dimensionality reduction on the input data. The obtained low-dimensional feature representation was spliced with the output of the mEE to obtain the final feature representation of the cell lines.
Finally, we fuse the feature representations of drugs and cell lines and use the fully connected layers to predict the drug response in cancer cell lines. We next describe the implementation of these channels.

Molecular fingerprint extractor and miRNA expression extractor based on 1D CNN

For input in conventional formats, such as the molecular fingerprints of drugs and miRNA expression of cell lines, we use a one-dimensional CNN to extract their features.
We use three convolutional layers in the model, with 4, 8, and 16 convolution kernels. Each element of a convolution kernel corresponds to a weight coefficient and a bias vector, similar to a neuron of a feedforward neural network [43]. Each neuron in a convolutional layer is connected to multiple neurons in a region close to the previous layer [44]. The size of the region depends on the size of the convolution kernel, which in our model is set to 8. This area is called a receptive field in the literature, whose meaning is analogous to that of a receptive field of a visual cortex cell [45]. When a convolution kernel is working, it scans the input features regularly, conducts matrix element multiplication and summation of input features in the receptive field, and superimposes the deviation [46], so as to achieve the effect of feature extraction,
$$\begin{aligned} &\mathbf {Z}^{l+1}(i) =\sum \limits _{k=1}^{K_{l}}\sum \limits _{x=1}^{f}[\mathbf {Z}_{k}^{l}(s_{0}i+x)\mathbf {w}_k^{l+1}(x)]+\mathbf {b},\\ i &\ \in \{0,1, \cdots ,L_{l+1}\}\quad L_{l+1}=\frac{L_{l}+2p-f}{s_{0}}+1. \end{aligned}$$
(1)
The summation in the formula is equivalent to solving a cross-correlation. \(\mathbf {b}\) is the amount of deviation, and \(\mathbf {Z}^{l}\) and \(\mathbf {Z}^{l+1}\) represent the input and output, respectively, of the \((l+1)\)th convolutional layer. \(L^{l+1}\) is the size of \(\mathbf {Z}^{l+1}\). The input is assumed to be one-dimensional, and convolved in one dimensional direction only, and the two-dimensional convolution formula [44] is similar to this. \(\mathbf {Z}(i)\) represents the values of the feature vector; K is the number of channels; and f, \(s_{0}\), and p are the parameters of the convolution layer, which represent the size of the convolution kernel, the stride, and the number of padding layers [46].
After feature extraction in each convolutional layer, the output feature data are passed to the pooling layer for feature selection and information filtering. The general form of \(L_{p}\) pooling is
$$\begin{aligned} \mathbf {A}_{k}^{l}(i)=\left[ \sum \limits _{x=1}^{f}\mathbf {A}_{k}^{l}(s_{0}i+x)^{p}\right] ^{\frac{1}{p}}, \end{aligned}$$
(2)
where p is a pre-specified parameter. When \(p=1\), \(L_{p}\) pooling takes the average value in the pooling area, which is called average pooling; when \(p\rightarrow \infty\), \(L_{p}\) pooling takes the maximum value in the area, i.e., max pooling [47]. Again, pooling is reduced to one-dimensional space. Our model uses the method of max pooling with a step size of 3, i.e., \(p\rightarrow \infty\), \(s_{0}=3\). It replaces the result of a single point in the feature vector with the feature statistics of its neighboring regions. After that, the features from the 16 channels are flattened into vectors, and the dimensions are converted to 128.

Copy number extractor based on stacked autoencoder

We cannot directly use conventional neural networks to extract features for DNA copy numbers with ultra-high-dimensions; we need to reduce the dimensionality in advance. Traditional methods such as PCA [48] can only reduce dimensionality in linear space and cannot perform nonlinear transformation, so we designed a stacked autoencoder [49] to predict the input by using fewer hidden nodes than the input nodes, i.e., to learn the function: \(h(x)\approx x\). In other words, it must learn an approximate identity function so that the output \(\hat{x}\) is approximately equal to the input x. For this reason, the network needs to encode as much information as possible into hidden nodes [50]. Stacked autoencoders are allowed to contain multiple hidden layers. We can learn more complex coding by adding hidden layers, but we must not make the autoencoder too powerful. If an encoder is too powerful, it just learns to map the input to an arbitrary number, and then the decoder learns its inverse mapping. Obviously, this autoencoder can reconstruct the data very well, but it cannot learn useful data representations. The autoencoder we designed contains six hidden layers, three belonging to the encoder and three to the decoder. The numbers of hidden layer neurons are 1024, 512, 256, 256, 512, and 1024. Because the traditional methods, such as the PCA method, can only reduce the dimensionality in linear space, we add nonlinear activation functions between the linear layers to enable nonlinear transformation. For the objective function during training, we use mean squared error, i.e.,
$$\begin{aligned} Loss=\frac{\sum \limits _{i=1}^{n} (\hat{y}_{i}-y_{i})^{2}}{n}, \end{aligned}$$
(3)
where y is the true value and \(\hat{y}\) is the predicted value. For ultra-high-dimensional and complex features of copy numbers, our model can encode these into low-dimensional data and represent the original feature well.

Drug structure extractor based on GCN

A CNN is only suitable for tensor data, such as two-dimensional images or one-dimensional text sequences. However, there is much data, whose relationships are difficult to simply express with tensors. For example, to use only a one-dimensional text sequence to represent the SMILES feature of a drug will lose its structural information. Thus, we need to use another common data structure, a graph represented by vertices and edges. Specifically, the SMILES sequence of a drug is transformed to the graph \(\mathbf {G}=(\mathbf {V},\mathbf {E})\) through RDKit and stored in the form of a feature matrix \(\mathbf {X}\) and an adjacency matrix \(\mathbf {A}\). \(\mathbf {X}\in \mathbf {R}^{n\times f}\) is composed of n nodes in the graph, and each node is represented by an f-dimensional vector. \(\mathbf {A}\in \mathbf {R}^{n\times n}\) represents an edge between nodes.
In order to extract the features of this kind of graph structure, we need to use a graph network. A currently popular method is to apply convolution to the graph structure, i.e., a GCN [51]. For the graph of SMILES, unlike matrix data, its convolution is difficult to define directly, so the convolution operation in the spatial domain must be transformed to matrix multiplication in the spectral domain,
$$\begin{aligned} \mathbf {g}_{\theta } *\mathbf {x}=\mathbf {U}\left( \mathbf {U}^{T}\mathbf {g}_{\theta } \cdot \mathbf {U}^{T}\mathbf {x}\right) , \end{aligned}$$
(4)
where g is the convolution kernel. The graph \(\mathbf {x}\) is represented as \(\mathbf {x}=(\mathbf {f}(1)\cdots \mathbf {f}(n))\in \mathbf {R}^{n}\), which is the signal at each point of the graph. \(\mathbf {U}\) is the basis of the Fourier transform and the eigenvector of the Laplacian matrix. However, the cost of calculating \(\mathbf {U}\) is too high, so after a series of approximate calculations, we obtain an approximate convolution formula,
$$\begin{aligned} \mathbf {g}_{\theta } *\mathbf {x}=\theta \left( \tilde{\mathbf {D}}^{-\frac{1}{2}} \tilde{\mathbf {A}}\tilde{\mathbf {D}}^{-\frac{1}{2}}\right) \mathbf {x}, \end{aligned}$$
(5)
where \(\tilde{\mathbf {A}}\) is the graph adjacency matrix with self-loop added, which sums the node itself when summing the eigenvectors of all adjacent nodes. It is thus possible to combine information of an atom in the drug compound with its neighbors. \(\tilde{\mathbf {D}}\) is the diagonal degree matrix of graph \(\tilde{\mathbf {A}}\), \(\tilde{\mathbf {D}}_{ii}=\sum \nolimits _{j} \tilde{\mathbf {A}}_{ij}\). The derivation process can be found in [51]. Then, after adding the nonlinear activation function \(\sigma\), we can train using the graph convolutional network,
$$\begin{aligned} \mathbf {H}^{(l+1)}=\sigma \left( \tilde{\mathbf {D}}^{-\frac{1}{2}} \tilde{\mathbf {A}}\tilde{\mathbf {D}}^{-\frac{1}{2}}\mathbf {H}^{(l)} \mathbf {W}^{(l)}\right) , \end{aligned}$$
(6)
where \(\mathbf {H}\) is the layer, and the superscript is the number of layers. Each additional graph convolution layer can aggregate the features of one more hop of neighbor nodes, thereby capturing as much neighborhood structure information as possible. \(\mathbf {H}^{(0)}\) is the feature matrix \(\mathbf {X}\), and \(\mathbf {W}\) is the trainable parameter matrix. We use three graph convolutional layers in the model, where the dimensions of \(\mathbf {W}^{(0)}\), \(\mathbf {W}^{(1)}\), and \(\mathbf {W}^{(2)}\) are \(f\times f\), \(f\times 2f\), and \(f\times 4f\), respectively. Thus, the dimensions of \(\mathbf {H}^{(1)}\), \(\mathbf {H}^{(2)}\), and \(\mathbf {H}^{(3)}\) are \(n\times f\), \(n\times 2f\), and \(n\times 4f\), respectively. We then use global maximum pooling to convert \(\mathbf {H}^{(3)}\) to a 4f-dimensional vector. Through the fully connected layers, the output dimension is 128.

Fusion layer

After the feature extraction channels, we concatenate the extracted features, fuse them through several fully connected layers, and make predictions. We add batch normalization (BN) layers between the linear layers and the nonlinear activation function to standardize the input of the activation function. This solves the problem of slow training due to inconsistent distributions of various features. Without normalization, the network needs more overhead to learn new distributions, which makes the model more complex and leads to overfitting. It also allows each layer to face the same distribution of input values, reducing the uncertainty caused by changes, and reducing the impact on subsequent layers. Each layer of the network becomes independent, which alleviates the problem of gradient disappearance in training.
After the sigmoid function, the output is mapped to (0, 1), which corresponds to the normalized value of a drug response. The steps of this method are shown as Algorithm 1.

Results

We divided drug-cell line pairs into training, validation, and test sets in an 8:1:1 ratio. The training set is used to train the models, and the model with the best result on the validation set is saved. We use the test set to test the saved model to obtain the final results. Further, we performed a five-fold cross-validation, that is, taking two pieces of data in turn as the validation set and the test set, and the remaining eight pieces as the training set. To evaluate these models, we use four classic metrics in regression: the Pearson correlation coefficient (\(CC_{p}\)), R squared (\(R^{2}\)), root mean square error (RMSE), and Spearman correlation coefficient (\(CC_{s}\)).
For NeRD, we adjusted hyperparameters such as dimensions after feature extraction, number of fusion layers, learning rate, epoch number, batch size, and dropout value according to the results of validation set. For those baseline methods, based on the principle of maintaining the original model, we also fine-tuned some hyperparameters according to the dataset we use to make the prediction results optimal. Details of hyperparameters are in Additional file 1: Table S2-S9.
After that, we designed six sets of experiments to verify the effectiveness of the proposed model from multiple perspectives.

Performance comparison

Our baseline includes classic machine learning methods—linear regression (LR) and random forest and support vector regression (SVR, SVR-L for linear kernel-based SVR); matrix factorization-based method—SRMF [52]; deep learning methods—MLP and CNN; and advanced dual-channel methods—VAE+MLP [53], tCNNS [33], CDRScan [31], DeepCDR [34], and GraphDRP [24]. We use the same data processing and division method to obtain experimental results through different models.
Table 1
Performance comparison. “\(\uparrow\)” means the larger the value, the better; “\(\downarrow\)” means the smaller the value, the better. The standard deviation of the cross-validation results is calculated by the STDEVP function
Method
\(CC_{p}\uparrow\)
\(R^{2}\uparrow\)
\(RMSE\downarrow\)
\(CC_{s}\uparrow\)
LR
0.234±0.0010
0.055±0.0004
0.171±0.0002
0.237±0.0011
SVR-L
0.232±0.0013
0.047±0.0012
0.172±0.0007
0.237±0.0008
SVR
0.469±0.0034
0.213±0.0065
0.153±0.0007
0.494±0.0051
RF
0.653±0.0355
0.419±0.0405
0.130±0.0056
0.606±0.0147
MLP
0.828±0.0029
0.698±0.0048
0.104±0.0007
0.800±0.0014
CNN
0.836±0.0026
0.700±0.0044
0.097±0.0008
0.807±0.0029
SRMF
0.837±0.0022
0.701±0.0037
0.097±0.0006
0.809±0.0018
VAE+MLP
0.830±0.0036
0.688±0.0060
0.098±0.0011
0.795±0.0031
DeepCDR
0.764±0.0147
0.572±0.0223
0.115±0.0032
0.676±0.0471
CDRScan
0.834±0.0039
0.696±0.0066
0.097±0.0008
0.810±0.0038
tCNNS
0.849±0.0039
0.721±0.0067
0.093±0.0010
0.822±0.0015
GraphDRP
0.848±0.0033
0.719±0.0057
0.093±0.0010
0.821±0.0020
NeRD
0.866\(\pm\)0.0027
0.750\(\pm\)0.0048
0.088\(\pm\)0.0007
0.839\(\pm\)0.0014
It can be seen from the results in Table 1 that our proposed model performs well, with a certain degree of improvement over each baseline. Our model shows an improvement of more than 4% over the best baseline on \(R^{2}\), and RMSE is reduced by 5% from the best baseline. \(CC_{p}\) and \(CC_{s}\) are also increased by more than 2%. It can also be seen from the results that the nonlinear regression method has an advantage on this problem, while the performance of the linear regression method is very poor.

Blind test

In performance comparison experiments, it may happen that the response data of a drug to some cell lines is divided into the training set, and the response data of this drug to other cell lines is divided into the test set. However, it may be necessary to predict the response of a new drug, and we designed a blind drug test for this purpose. We randomly select 10% of the drugs and use all drug-cell line pairs associated with them as the test set. Of the remaining 90% of drugs, 80% are used for training the model, and 10% for validation. It can also be necessary to predict the response of a new cell line, for which we designed a blind cell line test. We randomly select 90% of the cell lines and use all associated drug-cell line pairs for training, and the remaining 10% for testing. The number of data instances corresponding to each data partition can be found in Additional file 1: Table S10.
Table 2
Cell-line blind test
Method
\(CC_{p}\uparrow\)
\(R^{2}\uparrow\)
\(RMSE\downarrow\)
\(CC_{s}\uparrow\)
LR
0.231±0.0040
0.053±0.0020
0.171±0.0004
0.233±0.0039
SVR-L
0.110±0.0663
0.045±0.0206
0.180±0.0018
0.106±0.0634
SVR
0.471±0.0168
0.218±0.0169
0.153±0.0034
0.496±0.0015
RF
0.677±0.0351
0.440±0.0506
0.141±0.0044
0.566±0.0344
MLP
0.804±0.0144
0.658±0.0244
0.110±0.0040
0.767±0.0120
CNN
0.819±0.0124
0.671±0.0208
0.101±0.0034
0.781±0.0133
SRMF
0.836±0.0091
0.699±0.0151
0.096±0.0024
0.808±0.0093
VAE+MLP
0.796±0.0108
0.623±0.0187
0.108±0.0026
0.755±0.0102
DeepCDR
0.714±0.0568
0.506±0.0771
0.123±0.0094
0.586±0.1103
CDRScan
0.815±0.0123
0.663±0.0202
0.102±0.0034
0.788±0.0107
tCNNS
0.826±0.0156
0.682±0.0264
0.099±0.0044
0.791±0.0143
GraphDRP
0.833±0.0140
0.693±0.0228
0.097±0.0039
0.801±0.0125
NeRD
0.838\(\pm\)0.0132
0.702\(\pm\)0.0229
0.096\(\pm\)0.0039
0.808\(\pm\)0.0114
Table 3
Drug blind test
Method
\(CC_{p}\uparrow\)
\(R^{2}\uparrow\)
\(RMSE\downarrow\)
\(CC_{s}\uparrow\)
LR
0.201±0.0244
−0.029±0.0863
0.180±0.0132
0.196±0.0351
SVR-L
0.109±0.0435
−0.208±0.2698
0.192±0.0237
0.111±0.0444
SVR
0.315±0.1793
−0.274±0.3725
0.206±0.0076
0.254±0.1311
RF
0.112±0.2737
−0.461±0.1934
0.236±0.0515
0.137±0.2424
MLP
0.261±0.0435
−0.074±0.0924
0.184±0.0135
0.202±0.0701
CNN
0.223±0.0527
0.018±0.0468
0.176±0.0113
0.169±0.0689
SRMF
0.093±0.0553
−0.006±0.0316
0.311±0.0620
0.098±0.0437
VAE+MLP
0.283±0.0242
−0.190±0.0854
0.190±0.0087
0.238±0.0347
DeepCDR
0.318±0.1403
0.010±0.1699
0.174±0.0177
0.254±0.0922
CDRScan
0.297±0.0418
0.049±0.0278
0.173±0.0098
0.229±0.0583
tCNNS
0.256±0.0261
−0.029±0.1123
0.180±0.0179
0.230±0.0335
GraphDRP
0.312±0.0926
0.067±0.0799
0.172±0.0120
0.272±0.0653
NeRD
0.370\(\pm\)0.0131
0.069\(\pm\)0.0454
0.168\(\pm\)0.0076
0.291\(\pm\)0.0426
From the results of the blind test (Tables 2 and 3), it can be seen that the results of the blind cell line test are slightly lower than those of the mixed test, and the gap between different methods is not so obvious. It is worth noting that SRMF [52], a matrix factorization-based method, has almost no performance loss in blind cell line test, compared to mixed test. As Chen et al. [54] stated, some non-deep learning methods may work better in blind testing scenarios. However, the results of the blind drug test are unsatisfactory. This is predictable, because different cell lines still have strong similarities, but different drugs are not so similar, as Liu et al. [33] says. Consequently, when a drug to be predicted does not appear in the training set, it is difficult for the models to effectively extract its features and make correct predictions. This is a common problem in existing research [24, 33, 55], and even then, NeRD still outperforms baseline models. Surprisingly, SRMF did not perform as well as in the literature [54] on drug blind test. Therefore, we compared the data from GDSC in the original study with ours, which can be found in Additional file 1: Table S11.
In addition, random partitioning of dataset may lead to uncertainty in the results on blind test. It is more convincing to use drugs or cell lines with different similarities as test sets. To do this, we grouped drugs and cell lines by their level of similarity across the dataset, respectively, and then used each group as a test set in turn.
The prediction results of blind cell line test were positively correlated with the similarity level of test sets. The higher the similarity of the test set, the more accurate the prediction result. Blind drug test did not reflect this pattern. And no matter the scenario, NeRD still outperforms other methods. Specific results can be found in Additional file 1: Table S12.

Feature ablation experiment

We use multiple features of drugs and cell lines from different sources. We conducted a feature ablation experiment to verify the validity of the selected features. Specifically, we remove one feature of a drug or cell line, or remove one feature of each. We observe the results under these conditions and analyze the effect of each channel on the model’s performance.
It can be seen from the results (Fig. 2) that when any feature is lost, each evaluation index will drop slightly. This confirms that every feature we choose is beneficial to the model. It is interesting that when the molecular fingerprint of a drug is not used, the loss of performance is the most obvious, which shows that this is indeed a good feature to represent the drug. An intuitive result is that to only use the molecular fingerprint as the feature of a drug is better than just using SMILES, but this phenomenon does not appear in the two features of the cell line.
To further investigate the influence of each channel on the prediction results, we calculate the Shapley value for the four channels, which is the sum of the marginal contributions of each channel to the outcome divided by the number of possible combinations:
$$\begin{aligned} \begin{aligned} \varphi _{i}(v) = \frac{\sum \nolimits _{R} [v(S)-v(S-\{i\})]}{n!}, \end{aligned} \end{aligned}$$
(7)
where R is the permutation of n channels for a total of n!. S is a permutation in R, v(S) is the prediction result when channel i is included, and \(v(S-\{i\})\) is the outcome before adding channel i. Specifically, we calculate the Shapley values of four channels based on the evaluation indicators \(CC_{p}\), \(CC_{s}\), RMSE, and \(R^{2}\) respectively, and present them in the form of percentages.
Table 4
Influence of channels
 
Drug
Cell-line
Channel
DSE
MFE
mEE
CNE
\(CC_{p}\)
22.8%
30.5%
24.4%
22.3%
\(R^{2}\)
21.4%
33.4%
23.8%
21.4%
RMSE
20.0%
32.9%
24.7%
22.4%
\(CC_{s}\)
22.6%
32.3%
24.3%
20.8%
As can be seen from Table 4, each feature we selected plays an integral role. Among them, the molecular fingerprint of the drug have the greatest impact on the results, exceeding 30%, which means that the molecular fingerprint of the drug may represent itself better than the molecular graph. The difference between the two features of cell lines is small, and the influence of miRNA is slightly larger, which also shows that the influencing factors of cancer are multi-faceted.

Segment verification

To verify the effectiveness of the feature extraction and feature fusion parts of the model, we use the t-SNE algorithm to visualize the features of each stage. We analyze the effect of the model by observing the distribution of samples at different stages. We randomly select 1000 drug-cell line pairs. Before the feature is input to extraction channels, we concatenate the initial features and use t-SNE to map them to a two-dimensional space to facilitate the visualization of the sample distribution. To analyze the distinguishing ability of the feature representation, we use the value of IC50 as the label of the drug-cell line pairs to color the t-SNE graph. Similarly, the features after the four extraction channels are concatenated and mapped to a two-dimensional space, visualized, and colored. Features that have passed through the fully connected network of fusion layers are also presented.
It can be seen from Fig. 3 that before the feature extraction channels, drug-cell line pairs with different IC50 values are mixed together, with no regularity (Fig. 3a). After feature extraction, the data distribution becomes regular. Samples with high and low IC50 values are divided into the two ends of the picture, but the boundaries between other samples are not obvious (Fig. 3b). After the fusion layers, samples of middle-level IC50 are no longer mixed together, and all drug-cell line pairs are distributed in a segmented band according to the IC50 value (Fig. 3c). Data with different IC50 values are divided into different intervals.

Data reduction experiment

Due to the scarcity of labels in actual application, the effect of many models is often much less than the experimental effect. Thus, we artificially reduce the amount of training data and observe the attenuation of the effects of each model. We randomly select a portion of each training set in five-fold cross-validation for training, and the proportion of this portion is reduced from \(\frac{1}{2}\) to \(\frac{1}{16}\). Then, we test NeRD and several baselines with good experimental results with different amounts of training data.
The results of the data reduction experiment are shown in Fig. 4. It can be seen from the line charts (a–d) that the performance of each model is lost as the amount of data decreases. However, the prediction results of our model are relatively stable. Even when the amount of data is reduced to \(\frac{1}{16}\) of the total, it maintains a \(CC_{p}\) above 0.8 and an RMSE below 0.10. The performance degradation of other models is more obvious. In particular, GraphDRP, although it shows excellent performance on the original data, has results that deteriorate significantly as the amount of data continues to decrease, which may be due to the complexity of the model. To more intuitively observe the results, we draw the box plots (e, f) representing the distribution of prediction errors, from which it can be seen that when the data are sufficient (e), the prediction error of NeRD is slightly less than that of other methods. However, when data are scarce, the prediction error of the comparison methods deteriorates severely, while the results of NeRD remain stable (f).

Pharmacogenomics analysis

We use the trained NeRD model to predict unknows drug-cell line pairs in PRISM database (approximately 19.5% of all pairs across 388 cancer cell lines and 1448 drugs). To verify whether the predicted results have biological and clinical significance, we sorted the newly predicted IC50 values from small to large and selected the top 1% drug-cell line pairs (altogether 2537 pairs across 383 cancer cell lines and 91 drugs) (Additional file 1: Table S13). Based on the value of IC50, we have reason to believe that these drugs have certain anticancer activity against different cancer cell lines. For this reason, we found the target genes of these 91 drugs (Additional file 1: Table S14) according to the target information of the drugs provided in PRISM database. Then, we performed two global enrichment analyses of these genes, including Gene Ontology (GO) biological process and KEGG pathway enrichment. According to the results, these genes are significantly enriched in 364 GO terms and 110 pathways (adjusted p-value< 0.001). The top 20 enrichment results are shown in Fig. 5. GO enrichment analysis demonstrates multiple cancer-related processes (Fig. 5a), such as ion channels and transport [56], phosphorylation of the amino acid [57, 58], and phagocytosis [59]; these processes are intimately linked to tumor progression, maintenance, and treatment. KEGG pathway enrichment analysis reveals multiple significant biological pathways (Fig. 5b), which are strongly associated with cancer. These enriched pathways including ErbB signaling pathway [60], EGFR tyrosine kinase inhibitor resistance [61], viral carcinogenesis [62], proteasome [63], and apoptosis [64], and most of them have proven to be effective therapies against cancer.
Table 5
Case studies. Three largest cancer tissues (i.e., lung, skin, and pancreas) were screened from the predicted top 1% drug-cell line pairs, and then some drug-cell line pairs were screened from these tissues. We found that the predicted results of these drug-cell lines were consistent with those reported in the existing literature (i.e., Study)
Cell line
Cancer tissue
Drug name
Predicted IC50
Study
NCIH2122
Lung
Dasatinib
0.093358595
[65]
RERFLCAI
Lung
Dasatinib
0.124904478
[66]
NCIH1650
Lung
Bortezomib
0.046070208
[67]
NCIH322
Lung
Bortezomib
0.046246483
[11]
NCIH522
Lung
Ganetespib
0.074710398
[68]
SKMEL5
Skin
NVP-AUY922
0.038356718
[69]
UACC62
Skin
Piperazine
0.107990561
[70]
A2058
Skin
Piperazine
0.125350529
[71]
UACC62
Skin
Trametinib
0.063189984
[72]
WM1799
Skin
Trametinib
0.130639812
[73]
ASPC1
Pancreas
Dasatinib
0.120664515
[74]
PANC1005
Pancreas
Docetaxel
0.038846238
[75]
PATU8902
Pancreas
Docetaxel
0.041285745
[76]
SW1990
Pancreas
Docetaxel
0.049888730
[77]
HPAC
Pancreas
Ganetespib
0.070858083
[78]
We then also categorized the predicted top 1% of drug-cell line pairs according to the tissue that the cell line belonged to Fig. 6, selecting the three most numerous cancer tissues (i.e., lung, skin, and pancreas) for analysis. Importantly, we found that the predictive results for many of these cell line drug pairs in these tissues have been confirmed by the existing literature (Table 5). For example, in the analysis of lung cancer, dasatinib as a Src family kinases (SFKs) inhibitor can inhibit the growth and survival of non-small cell lung cancer NCI-H2122 cells [65]. In skin cancer, NVP-AUY922, a heat shock protein 90 (HSP90) inhibitor can sensitize melanoma SKMEL5 cells to it [69]. In pancreatic cancer studies, pancreatic cancer PANC1005 cells are sensitive to the tubulin polymerization inhibitor docetaxel, which is consistent with our predicted results [75]. Taken together, these case studies support that NeRD is able to effectively predict the drug sensitivity of cell lines, which can help speed up the screening of drugs and find new anti-cancer drugs in actual clinical settings.

Discussion

We presented a multichannel neural network model, NeRD, to computationally predict cancer drug responses by integrating multi-dimensional data. We designed feature extractors DSE, MFE, mEE, and CNE to extract informative embeddings from multidimensional features of cell lines and drugs. Features extracted from each channel were converted to a uniform format, fused, and predicted. The results of five experiments show that NeRD achieves excellent performance from many aspects. First, it performs better than comparative models. Second, its generalizability was demonstrated by blind test results, and it outperformed other models when predicting new samples. Third, the results of a feature ablation experiment show that each selected feature is beneficial to the model, and that NeRD effectively fuses multiple information sources and features from different data structures and dimensions. Fourth, according to a segment verification experiment, NeRD has a strong feature extraction capability, which indirectly shows that each feature extractor designed in the model has strong utility. Fifth, NeRD has high robustness, as illustrated by a data reduction experiment. Sixth, the result of using trained NeRD for drug sensitivity prediction have biological and clinical significance. Despite NeRD having strong predictive power, the model was built on in vitro data. Challenges remain in its application. Recent studies have shown that using clinical data from some patients can better help achieve precision oncology [27, 79]. These challenges can be addressed in our future studies.

Conclusion

In summary, we think that NeRD, as a highly extensible framework, can effectively fuse multidimensional features of cell lines and drugs to accurately predict the drug response of cell lines. Furthermore, this model can be widely applicable to integrate other omics data, thus benefiting clinical cancer therapy and future research on drug response prediction. Thus, it will provide a more diverse view of clinical cancer therapy.

Acknowledgements

Not applicable.

Declarations

Not applicable.
Not applicable.

Competing interests

The authors declare that they have no competing interests.
Open AccessThis 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.

Publisher’s Note

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

Supplementary information

Additional file 1: Table S1. RDKit functions and their descriptions. Table S2. Hyperparameters for NeRD. The adjustment of hyperparameters often has an important impact on the specific data set. Table S3. Hyperparameters for DeepCDR, CDRScan, tCNNS, and GraphDRP. These models are all dual-channel or quasi-dual-channel, so the same method is used to adjust the hyperparameters. Table S4. Hyperparameters for RF. The parameters of the RF framework are few, and the parameter selection is generally to adjust the value of N\_estimators, i.e., the number of decision trees. Table S5. Hyperparameters for SVR. Gamma is the coefficient of kernel functions, only valid for `rbf', `poly', and `sigmod'. The parameter Degree only works for `kernel=poly'. C represents the penalty coefficient of the error term. The larger C is, the greater the degree of penalty for wrongly classified samples. Table S6. Hyperparameters for CNN. What we use here is the one-dimensional convolution function provided by pytorch. Table S7. Hyperparameters for MLP. The number of neurons in each layer is also fine-tuned according to the number of hidden layers. Table S8. Hyperparameters for SRMF. SRMF is a method based on matrix factorization, and its hyperparameters mainly include the dimension of the feature space and the regularization parameters. Table S9. Hyperparameters for VAE+MLP. The number of neurons in each layer is also fine-tuned according to the number of hidden layers. Table S10. Number of data instances corresponds to each data partition in the blind test. Table S11. Dataset comparison. Table S12. Blind test dividing data by similarity. Set1-Set5 are test sets with increasing similarity. Set1 has the lowest similarity and Set5 has the highest similarity. The values are the Pearson correlation coefficients. Table S13. Predicted results for the top 1\% of drug-cell lines. We used the trained NERD model to predict drug cell line pairs without IC50 data in the PRISM database, sorted from small to large according to the predicted IC50 value, and then screened the top 1\% of drug-cell line pairs (altogether 2537 pairs across 383 cancer cell lines and 91 drugs). Table S14. The drug target gene list. Based on the list of drugs obtained from the top 1\% of predicted drug-cell line pairs, we found the target genes for these drugs from the PRISM database.
Literatur
1.
Zurück zum Zitat Dagogo-Jack I, Shaw AT. Tumour heterogeneity and resistance to cancer therapies. Nat Rev Clin Oncol. 2018;15(2):81–94.PubMedCrossRef Dagogo-Jack I, Shaw AT. Tumour heterogeneity and resistance to cancer therapies. Nat Rev Clin Oncol. 2018;15(2):81–94.PubMedCrossRef
3.
Zurück zum Zitat Azuaje F. Computational models for predicting drug responses in cancer research. Brief Bioinform. 2017;18(5):820–9. Azuaje F. Computational models for predicting drug responses in cancer research. Brief Bioinform. 2017;18(5):820–9.
4.
Zurück zum Zitat Baptista D, Ferreira PG, Rocha M. Deep learning for drug response prediction in cancer. Brief Bioinform. 2021;22(1):360–79.PubMedCrossRef Baptista D, Ferreira PG, Rocha M. Deep learning for drug response prediction in cancer. Brief Bioinform. 2021;22(1):360–79.PubMedCrossRef
5.
Zurück zum Zitat Garraway LA, Verweij J, Ballman KV, et al. Precision oncology: an overview. J Clin Oncol. 2013;31(15):1803–5.PubMedCrossRef Garraway LA, Verweij J, Ballman KV, et al. Precision oncology: an overview. J Clin Oncol. 2013;31(15):1803–5.PubMedCrossRef
8.
Zurück zum Zitat Rahman R, Dhruba SR, Matlock K, De-Niz C, Ghosh S, Pal R. Evaluating the consistency of large-scale pharmacogenomic studies. Brief Bioinform. 2019;20(5):1734–53.PubMedPubMedCentralCrossRef Rahman R, Dhruba SR, Matlock K, De-Niz C, Ghosh S, Pal R. Evaluating the consistency of large-scale pharmacogenomic studies. Brief Bioinform. 2019;20(5):1734–53.PubMedPubMedCentralCrossRef
9.
Zurück zum Zitat Yang W, Soares J, Greninger P, Edelman EJ, Lightfoot H, Forbes S, et al. Genomics of Drug Sensitivity in Cancer (GDSC): a resource for therapeutic biomarker discovery in cancer cells. Nucleic Acids Res. 2012;41(D1):D955–61.PubMedPubMedCentralCrossRef Yang W, Soares J, Greninger P, Edelman EJ, Lightfoot H, Forbes S, et al. Genomics of Drug Sensitivity in Cancer (GDSC): a resource for therapeutic biomarker discovery in cancer cells. Nucleic Acids Res. 2012;41(D1):D955–61.PubMedPubMedCentralCrossRef
11.
Zurück zum Zitat Shoemaker RH. The NCI60 human tumour cell line anticancer drug screen. Nat Rev Cancer. 2006;6(10):813–23.PubMedCrossRef Shoemaker RH. The NCI60 human tumour cell line anticancer drug screen. Nat Rev Cancer. 2006;6(10):813–23.PubMedCrossRef
13.
Zurück zum Zitat Yu C, Mannan AM, Yvone GM, Ross KN, Zhang YL, Marton MA, et al. High-throughput identification of genotype-specific cancer vulnerabilities in mixtures of barcoded tumor cell lines. Nat Biotechnol. 2016;34(4):419–23.PubMedPubMedCentralCrossRef Yu C, Mannan AM, Yvone GM, Ross KN, Zhang YL, Marton MA, et al. High-throughput identification of genotype-specific cancer vulnerabilities in mixtures of barcoded tumor cell lines. Nat Biotechnol. 2016;34(4):419–23.PubMedPubMedCentralCrossRef
14.
15.
Zurück zum Zitat Menden MP, Iorio F, Garnett M, McDermott U, Benes CH, Ballester PJ, et al. Machine learning prediction of cancer cell sensitivity to drugs based on genomic and chemical properties. PLoS ONE. 2013;8(4): e61318.PubMedPubMedCentralCrossRef Menden MP, Iorio F, Garnett M, McDermott U, Benes CH, Ballester PJ, et al. Machine learning prediction of cancer cell sensitivity to drugs based on genomic and chemical properties. PLoS ONE. 2013;8(4): e61318.PubMedPubMedCentralCrossRef
16.
Zurück zum Zitat Liu C, Wei D, Xiang J, Ren F, Huang L, Lang J, et al. An improved anticancer drug-response prediction based on an ensemble method integrating matrix completion and ridge regression. Mol Ther Nucleic Acids. 2020;21:676–86.PubMedPubMedCentralCrossRef Liu C, Wei D, Xiang J, Ren F, Huang L, Lang J, et al. An improved anticancer drug-response prediction based on an ensemble method integrating matrix completion and ridge regression. Mol Ther Nucleic Acids. 2020;21:676–86.PubMedPubMedCentralCrossRef
17.
Zurück zum Zitat Huang EW, Bhope A, Lim J, Sinha S, Emad A. Tissue-guided LASSO for prediction of clinical drug response using preclinical samples. PLoS Comput Biol. 2020;16(1): e1007607.PubMedPubMedCentralCrossRef Huang EW, Bhope A, Lim J, Sinha S, Emad A. Tissue-guided LASSO for prediction of clinical drug response using preclinical samples. PLoS Comput Biol. 2020;16(1): e1007607.PubMedPubMedCentralCrossRef
18.
Zurück zum Zitat Clayton EA, Pujol TA, McDonald JF, Qiu P. Leveraging TCGA gene expression data to build predictive models for cancer drug response. BMC Bioinforma. 2020;21(14):1–11. Clayton EA, Pujol TA, McDonald JF, Qiu P. Leveraging TCGA gene expression data to build predictive models for cancer drug response. BMC Bioinforma. 2020;21(14):1–11.
19.
Zurück zum Zitat Costello JC, Heiser LM, Georgii E, Gönen M, Menden MP, Wang NJ, et al. A community effort to assess and improve drug sensitivity prediction algorithms. Nat Biotechnol. 2014;32(12):1202–12.PubMedPubMedCentralCrossRef Costello JC, Heiser LM, Georgii E, Gönen M, Menden MP, Wang NJ, et al. A community effort to assess and improve drug sensitivity prediction algorithms. Nat Biotechnol. 2014;32(12):1202–12.PubMedPubMedCentralCrossRef
20.
Zurück zum Zitat Ammad-Ud-Din M, Khan SA, Malani D, Murumägi A, Kallioniemi O, Aittokallio T, et al. Drug response prediction by inferring pathway-response associations with kernelized Bayesian matrix factorization. Bioinformatics. 2016;32(17):i455-63.PubMedCrossRef Ammad-Ud-Din M, Khan SA, Malani D, Murumägi A, Kallioniemi O, Aittokallio T, et al. Drug response prediction by inferring pathway-response associations with kernelized Bayesian matrix factorization. Bioinformatics. 2016;32(17):i455-63.PubMedCrossRef
21.
Zurück zum Zitat Berlow N, Haider S, Wan Q, Geltzeiler M, Davis LE, Keller C, et al. An integrated approach to anti-cancer drug sensitivity prediction. IEEE/ACM Trans Comput Biol Bioinform. 2014;11(6):995–1008.PubMedCrossRef Berlow N, Haider S, Wan Q, Geltzeiler M, Davis LE, Keller C, et al. An integrated approach to anti-cancer drug sensitivity prediction. IEEE/ACM Trans Comput Biol Bioinform. 2014;11(6):995–1008.PubMedCrossRef
22.
Zurück zum Zitat Yuan H, Paskov I, Paskov H, González AJ, Leslie CS. Multitask learning improves prediction of cancer drug sensitivity. Sci Rep. 2016;6(1):1–11. Yuan H, Paskov I, Paskov H, González AJ, Leslie CS. Multitask learning improves prediction of cancer drug sensitivity. Sci Rep. 2016;6(1):1–11.
23.
Zurück zum Zitat Sharma A, Rani R. Drug sensitivity prediction framework using ensemble and multi-task learning. Int J Mach Learn Cybern. 2020;11(6):1231–40.CrossRef Sharma A, Rani R. Drug sensitivity prediction framework using ensemble and multi-task learning. Int J Mach Learn Cybern. 2020;11(6):1231–40.CrossRef
24.
Zurück zum Zitat Nguyen TT, Nguyen GTT, Nguyen T, Le DH. Graph convolutional networks for drug response prediction. IEEE/ACM Trans Comput Biol Bioinform. 2021;19(1):146–54. Nguyen TT, Nguyen GTT, Nguyen T, Le DH. Graph convolutional networks for drug response prediction. IEEE/ACM Trans Comput Biol Bioinform. 2021;19(1):146–54.
25.
Zurück zum Zitat Parca L, Pepe G, Pietrosanto M, Galvan G, Galli L, Palmeri A, et al. Modeling cancer drug response through drug-specific informative genes. Sci Rep. 2019;9(1):1–11.CrossRef Parca L, Pepe G, Pietrosanto M, Galvan G, Galli L, Palmeri A, et al. Modeling cancer drug response through drug-specific informative genes. Sci Rep. 2019;9(1):1–11.CrossRef
26.
Zurück zum Zitat Naulaerts S, Menden MP, Ballester PJ. Concise polygenic models for cancer-specific identification of drug-sensitive tumors from their multi-omics profiles. Biomolecules. 2020;10(6):963.PubMedCentralCrossRef Naulaerts S, Menden MP, Ballester PJ. Concise polygenic models for cancer-specific identification of drug-sensitive tumors from their multi-omics profiles. Biomolecules. 2020;10(6):963.PubMedCentralCrossRef
27.
Zurück zum Zitat Huang C, Clayton EA, Matyunina LV, McDonald L, Benigno BB, Vannberg F, et al. Machine learning predicts individual cancer patient responses to therapeutic drugs with high accuracy. Sci Rep. 2018;8(1):1–8.CrossRef Huang C, Clayton EA, Matyunina LV, McDonald L, Benigno BB, Vannberg F, et al. Machine learning predicts individual cancer patient responses to therapeutic drugs with high accuracy. Sci Rep. 2018;8(1):1–8.CrossRef
28.
Zurück zum Zitat Firoozbakht F, Yousefi B, Schwikowski B. An overview of machine learning methods for monotherapy drug response prediction. Brief Bioinform. 2022;23(1):bbab408.PubMedCrossRef Firoozbakht F, Yousefi B, Schwikowski B. An overview of machine learning methods for monotherapy drug response prediction. Brief Bioinform. 2022;23(1):bbab408.PubMedCrossRef
29.
Zurück zum Zitat Emdadi A, Eslahchi C. Auto-HMM-LMF: feature selection based method for prediction of drug response via autoencoder and hidden Markov model. BMC Bioinformatics. 2021;22(1):1–22.CrossRef Emdadi A, Eslahchi C. Auto-HMM-LMF: feature selection based method for prediction of drug response via autoencoder and hidden Markov model. BMC Bioinformatics. 2021;22(1):1–22.CrossRef
30.
Zurück zum Zitat Rampášek L, Hidru D, Smirnov P, Haibe-Kains B, Goldenberg A. Dr. VAE: improving drug response prediction via modeling of drug perturbation effects. Bioinformatics. 2019;35(19):3743–51.PubMedPubMedCentralCrossRef Rampášek L, Hidru D, Smirnov P, Haibe-Kains B, Goldenberg A. Dr. VAE: improving drug response prediction via modeling of drug perturbation effects. Bioinformatics. 2019;35(19):3743–51.PubMedPubMedCentralCrossRef
31.
Zurück zum Zitat Chang Y, Park H, Yang HJ, Lee S, Lee KY, Kim TS, et al. Cancer drug response profile scan (CDRscan): a deep learning model that predicts drug effectiveness from cancer genomic signature. Sci Rep. 2018;8(1):1–11. Chang Y, Park H, Yang HJ, Lee S, Lee KY, Kim TS, et al. Cancer drug response profile scan (CDRscan): a deep learning model that predicts drug effectiveness from cancer genomic signature. Sci Rep. 2018;8(1):1–11.
32.
Zurück zum Zitat Su R, Liu X, Xiao G, Wei L. Meta-GDBP: a high-level stacked regression model to improve anticancer drug response prediction. Brief Bioinform. 2020;21(3):996–1005.PubMedCrossRef Su R, Liu X, Xiao G, Wei L. Meta-GDBP: a high-level stacked regression model to improve anticancer drug response prediction. Brief Bioinform. 2020;21(3):996–1005.PubMedCrossRef
33.
Zurück zum Zitat Liu P, Li H, Li S, Leung KS. Improving prediction of phenotypic drug response on cancer cell lines using deep convolutional network. BMC Bioinforma. 2019;20(1):1–14.CrossRef Liu P, Li H, Li S, Leung KS. Improving prediction of phenotypic drug response on cancer cell lines using deep convolutional network. BMC Bioinforma. 2019;20(1):1–14.CrossRef
34.
Zurück zum Zitat Liu Q, Hu Z, Jiang R, Zhou M. DeepCDR: a hybrid graph convolutional network for predicting cancer drug response. Bioinformatics. 2020;36(Supplement-2):i911–8.PubMedCrossRef Liu Q, Hu Z, Jiang R, Zhou M. DeepCDR: a hybrid graph convolutional network for predicting cancer drug response. Bioinformatics. 2020;36(Supplement-2):i911–8.PubMedCrossRef
35.
Zurück zum Zitat Landrum G. RDKit documentation. Release. 2013;1:1–79. Landrum G. RDKit documentation. Release. 2013;1:1–79.
37.
39.
Zurück zum Zitat Solomon DA, Kim JS, Ressom HW, Sibenaller Z, Ryken T, Jean W, et al. Sample type bias in the analysis of cancer genomes. Cancer Res. 2009;69(14):5630–3.PubMedPubMedCentralCrossRef Solomon DA, Kim JS, Ressom HW, Sibenaller Z, Ryken T, Jean W, et al. Sample type bias in the analysis of cancer genomes. Cancer Res. 2009;69(14):5630–3.PubMedPubMedCentralCrossRef
40.
Zurück zum Zitat Bartel DP. MicroRNAs: genomics, biogenesis, mechanism, and function. Cell. 2004;116(2):281–97.PubMedCrossRef Bartel DP. MicroRNAs: genomics, biogenesis, mechanism, and function. Cell. 2004;116(2):281–97.PubMedCrossRef
41.
Zurück zum Zitat Rupaimoole R, Slack FJ. MicroRNA therapeutics: towards a new era for the management of cancer and other diseases. Nat Rev Drug Discov. 2017;16(3):203–22.PubMedCrossRef Rupaimoole R, Slack FJ. MicroRNA therapeutics: towards a new era for the management of cancer and other diseases. Nat Rev Drug Discov. 2017;16(3):203–22.PubMedCrossRef
42.
Zurück zum Zitat Jolly S, Lang V, Koelzer VH, Frigerio CS, Magno L, Salinas PC, et al. Single-cell quantification of mRNA expression in the human brain. Sci Rep. 2019;9(1):1–9.CrossRef Jolly S, Lang V, Koelzer VH, Frigerio CS, Magno L, Salinas PC, et al. Single-cell quantification of mRNA expression in the human brain. Sci Rep. 2019;9(1):1–9.CrossRef
43.
Zurück zum Zitat Qin C, Shi G, Tao J, Yu H, Jin Y, Lei J, et al. Precise cutterhead torque prediction for shield tunneling machines using a novel hybrid deep neural network. Mech Syst Signal Process. 2021;151:107386.CrossRef Qin C, Shi G, Tao J, Yu H, Jin Y, Lei J, et al. Precise cutterhead torque prediction for shield tunneling machines using a novel hybrid deep neural network. Mech Syst Signal Process. 2021;151:107386.CrossRef
44.
Zurück zum Zitat Albawi S, Mohammed TA, Al-Zawi S. Understanding of a convolutional neural network. In: 2017 International Conference on Engineering and Technology (ICET). Piscataway: IEEE; 2017. p. 1–6. Albawi S, Mohammed TA, Al-Zawi S. Understanding of a convolutional neural network. In: 2017 International Conference on Engineering and Technology (ICET). Piscataway: IEEE; 2017. p. 1–6.
45.
Zurück zum Zitat Gu J, Wang Z, Kuen J, Ma L, Shahroudy A, Shuai B, et al. Recent advances in convolutional neural networks. Pattern Recognit. 2018;77:354–77.CrossRef Gu J, Wang Z, Kuen J, Ma L, Shahroudy A, Shuai B, et al. Recent advances in convolutional neural networks. Pattern Recognit. 2018;77:354–77.CrossRef
46.
Zurück zum Zitat Goodfellow I, Bengio Y, Courville A. Deep learning. Cambridge: MIT press; 2016. Goodfellow I, Bengio Y, Courville A. Deep learning. Cambridge: MIT press; 2016.
47.
Zurück zum Zitat Akhtar N, Ragavendran U. Interpretation of intelligence in CNN-pooling processes: a methodological survey. Neural Comput Appl. 2020;32(3):879–98.CrossRef Akhtar N, Ragavendran U. Interpretation of intelligence in CNN-pooling processes: a methodological survey. Neural Comput Appl. 2020;32(3):879–98.CrossRef
48.
Zurück zum Zitat Abdi H, Williams LJ. Principal component analysis. Wiley Interdiscip Rev Comput Stat. 2010;2(4):433–59.CrossRef Abdi H, Williams LJ. Principal component analysis. Wiley Interdiscip Rev Comput Stat. 2010;2(4):433–59.CrossRef
49.
Zurück zum Zitat Zabalza J, Ren J, Zheng J, Zhao H, Qing C, Yang Z, et al. Novel segmented stacked autoencoder for effective dimensionality reduction and feature extraction in hyperspectral imaging. Neurocomputing. 2016;185:1–10.CrossRef Zabalza J, Ren J, Zheng J, Zhao H, Qing C, Yang Z, et al. Novel segmented stacked autoencoder for effective dimensionality reduction and feature extraction in hyperspectral imaging. Neurocomputing. 2016;185:1–10.CrossRef
50.
Zurück zum Zitat Xie R, Wen J, Quitadamo A, Cheng J, Shi X. A deep auto-encoder model for gene expression prediction. BMC Genom. 2017;18(9):39–49. Xie R, Wen J, Quitadamo A, Cheng J, Shi X. A deep auto-encoder model for gene expression prediction. BMC Genom. 2017;18(9):39–49.
51.
Zurück zum Zitat Kipf TN, Welling M. Semi-Supervised Classification with Graph Convolutional Networks. arXiv e-prints. 2016;p. arXiv:1609.02907. Kipf TN, Welling M. Semi-Supervised Classification with Graph Convolutional Networks. arXiv e-prints. 2016;p. arXiv:1609.02907.
52.
Zurück zum Zitat Wang L, Li X, Zhang L, Gao Q. Improved anticancer drug response prediction in cell lines using matrix factorization with similarity regularization. BMC Cancer. 2017;17(1):1–12.CrossRef Wang L, Li X, Zhang L, Gao Q. Improved anticancer drug response prediction in cell lines using matrix factorization with similarity regularization. BMC Cancer. 2017;17(1):1–12.CrossRef
53.
Zurück zum Zitat Dong H, Xie J, Jing Z, Ren D. Variational Autoencoder for Anti-Cancer Drug Response Prediction. arXiv e-prints. 2020;p.arXiv:2008.09763. Dong H, Xie J, Jing Z, Ren D. Variational Autoencoder for Anti-Cancer Drug Response Prediction. arXiv e-prints. 2020;p.arXiv:2008.09763.
54.
Zurück zum Zitat Chen Y, Zhang L. How much can deep learning improve prediction of the responses to drugs in cancer cell lines? Brief Bioinform. 2022;23(1):bbab378. Chen Y, Zhang L. How much can deep learning improve prediction of the responses to drugs in cancer cell lines? Brief Bioinform. 2022;23(1):bbab378.
55.
Zurück zum Zitat Zhu Y, Brettin T, Evrard YA, Partin A, Xia F, Shukla M, et al. Ensemble transfer learning for the prediction of anti-cancer drug response. Sci Rep. 2020;10(1):1–11.CrossRef Zhu Y, Brettin T, Evrard YA, Partin A, Xia F, Shukla M, et al. Ensemble transfer learning for the prediction of anti-cancer drug response. Sci Rep. 2020;10(1):1–11.CrossRef
56.
Zurück zum Zitat Pedersen SF, Stock C. Ion channels and transporters in cancer: pathophysiology, regulation, and clinical potential. Cancer Res. 2013;73(6):1658–61.PubMedCrossRef Pedersen SF, Stock C. Ion channels and transporters in cancer: pathophysiology, regulation, and clinical potential. Cancer Res. 2013;73(6):1658–61.PubMedCrossRef
58.
Zurück zum Zitat Julien SG, Dubé N, Hardy S, Tremblay ML. Inside the human cancer tyrosine phosphatome. Nat Rev Cancer. 2011;11(1):35–49.PubMedCrossRef Julien SG, Dubé N, Hardy S, Tremblay ML. Inside the human cancer tyrosine phosphatome. Nat Rev Cancer. 2011;11(1):35–49.PubMedCrossRef
59.
Zurück zum Zitat Feng M, Jiang W, Kim B, Zhang CC, Fu YX, Weissman IL. Phagocytosis checkpoints as new targets for cancer immunotherapy. Nat Rev Cancer. 2019;19(10):568–86.PubMedPubMedCentralCrossRef Feng M, Jiang W, Kim B, Zhang CC, Fu YX, Weissman IL. Phagocytosis checkpoints as new targets for cancer immunotherapy. Nat Rev Cancer. 2019;19(10):568–86.PubMedPubMedCentralCrossRef
60.
Zurück zum Zitat Kumagai S, Koyama S, Nishikawa H. Antitumour immunity regulated by aberrant ERBB family signalling. Nat Rev Cancer. 2021;21(3):181–97.PubMedCrossRef Kumagai S, Koyama S, Nishikawa H. Antitumour immunity regulated by aberrant ERBB family signalling. Nat Rev Cancer. 2021;21(3):181–97.PubMedCrossRef
63.
Zurück zum Zitat Rousseau A, Bertolotti A. Regulation of proteasome assembly and activity in health and disease. Nat Rev Mol Cell Biol. 2018;19(11):697–712.PubMedCrossRef Rousseau A, Bertolotti A. Regulation of proteasome assembly and activity in health and disease. Nat Rev Mol Cell Biol. 2018;19(11):697–712.PubMedCrossRef
64.
Zurück zum Zitat Wong RS. Apoptosis in cancer: from pathogenesis to treatment. J Exp Clin Cancer Res. 2011;30(1):1–14.CrossRef Wong RS. Apoptosis in cancer: from pathogenesis to treatment. J Exp Clin Cancer Res. 2011;30(1):1–14.CrossRef
65.
Zurück zum Zitat Leroy C, Shen Q, Strande V, Meyer R, McLaughlin M, Lezan E, et al. CUB-domain-containing protein 1 overexpression in solid cancers promotes cancer cell growth by activating Src family kinases. Oncogene. 2015;34(44):5593–8.PubMedPubMedCentralCrossRef Leroy C, Shen Q, Strande V, Meyer R, McLaughlin M, Lezan E, et al. CUB-domain-containing protein 1 overexpression in solid cancers promotes cancer cell growth by activating Src family kinases. Oncogene. 2015;34(44):5593–8.PubMedPubMedCentralCrossRef
66.
Zurück zum Zitat Hamanaka N, Nakanishi Y, Mizuno T, Horiguchi-Takei K, Akiyama N, Tanimura H, et al. YES1 is a targetable oncogene in cancers harboring YES1 gene amplification. Cancer Res. 2019;79(22):5734–45.PubMedCrossRef Hamanaka N, Nakanishi Y, Mizuno T, Horiguchi-Takei K, Akiyama N, Tanimura H, et al. YES1 is a targetable oncogene in cancers harboring YES1 gene amplification. Cancer Res. 2019;79(22):5734–45.PubMedCrossRef
67.
Zurück zum Zitat Wang WH, Zhan JM, Tang YL, Zhou N, Liu WY, Jiang DW. miR-466 contributes to the enhanced antitumor effect of bortezomib on non-small-cell lung cancer by inhibiting CCND1. Chemotherapy. 2022;67(2):110–22. Wang WH, Zhan JM, Tang YL, Zhou N, Liu WY, Jiang DW. miR-466 contributes to the enhanced antitumor effect of bortezomib on non-small-cell lung cancer by inhibiting CCND1. Chemotherapy. 2022;67(2):110–22.
68.
Zurück zum Zitat Shimamura T, Perera SA, Foley KP, Sang J, Rodig SJ, Inoue T, et al. Ganetespib (STA-9090), a nongeldanamycin HSP90 inhibitor, has potent antitumor activity in in vitro and in vivo models of non-small cell lung cancer. Clin Cancer Res. 2012;18(18):4973–85.PubMedPubMedCentralCrossRef Shimamura T, Perera SA, Foley KP, Sang J, Rodig SJ, Inoue T, et al. Ganetespib (STA-9090), a nongeldanamycin HSP90 inhibitor, has potent antitumor activity in in vitro and in vivo models of non-small cell lung cancer. Clin Cancer Res. 2012;18(18):4973–85.PubMedPubMedCentralCrossRef
69.
Zurück zum Zitat Eccles SA, Massey A, Raynaud FI, Sharp SY, Box G, Valenti M, et al. NVP-AUY922: a novel heat shock protein 90 inhibitor active against xenograft tumor growth, angiogenesis, and metastasis. Cancer Res. 2008;68(8):2850–60.PubMedCrossRef Eccles SA, Massey A, Raynaud FI, Sharp SY, Box G, Valenti M, et al. NVP-AUY922: a novel heat shock protein 90 inhibitor active against xenograft tumor growth, angiogenesis, and metastasis. Cancer Res. 2008;68(8):2850–60.PubMedCrossRef
70.
Zurück zum Zitat Wong HN, Lewies A, Haigh M, Viljoen JM, Wentzel JF, Haynes RK, et al. Anti-melanoma activities of artemisone and prenylated amino-artemisinins in combination with known anticancer drugs. Front Pharmacol. 2020;11:1543. Wong HN, Lewies A, Haigh M, Viljoen JM, Wentzel JF, Haynes RK, et al. Anti-melanoma activities of artemisone and prenylated amino-artemisinins in combination with known anticancer drugs. Front Pharmacol. 2020;11:1543.
71.
Zurück zum Zitat Maquoi E, Sounni NE, Devy L, Olivier F, Frankenne F, Krell HW, et al. Anti-invasive, antitumoral, and antiangiogenic efficacy of a pyrimidine-2, 4, 6-trione derivative, an orally active and selective matrix metalloproteinases inhibitor. Clin Cancer Res. 2004;10(12):4038–47.PubMedCrossRef Maquoi E, Sounni NE, Devy L, Olivier F, Frankenne F, Krell HW, et al. Anti-invasive, antitumoral, and antiangiogenic efficacy of a pyrimidine-2, 4, 6-trione derivative, an orally active and selective matrix metalloproteinases inhibitor. Clin Cancer Res. 2004;10(12):4038–47.PubMedCrossRef
72.
Zurück zum Zitat Cohen-Solal KA, Kaufman HL, Lasfar A. Transcription factors as critical players in melanoma invasiveness, drug resistance, and opportunities for therapeutic drug development. Pigment Cell Melanoma Res. 2018;31(2):241–52.PubMedCrossRef Cohen-Solal KA, Kaufman HL, Lasfar A. Transcription factors as critical players in melanoma invasiveness, drug resistance, and opportunities for therapeutic drug development. Pigment Cell Melanoma Res. 2018;31(2):241–52.PubMedCrossRef
73.
Zurück zum Zitat Gopal YV, Gammon S, Prasad R, Knighton B, Pisaneschi F, Roszik J, et al. A novel mitochondrial inhibitor blocks MAPK pathway and overcomes MAPK inhibitor resistance in melanoma. Clin Cancer Res. 2019;25(21):6429–42.CrossRef Gopal YV, Gammon S, Prasad R, Knighton B, Pisaneschi F, Roszik J, et al. A novel mitochondrial inhibitor blocks MAPK pathway and overcomes MAPK inhibitor resistance in melanoma. Clin Cancer Res. 2019;25(21):6429–42.CrossRef
74.
Zurück zum Zitat Ma L, Wei J, Su GH, Lin J. Dasatinib can enhance paclitaxel and gemcitabine inhibitory activity in human pancreatic cancer cells. Cancer Biol Ther. 2019;20(6):855–65.PubMedPubMedCentralCrossRef Ma L, Wei J, Su GH, Lin J. Dasatinib can enhance paclitaxel and gemcitabine inhibitory activity in human pancreatic cancer cells. Cancer Biol Ther. 2019;20(6):855–65.PubMedPubMedCentralCrossRef
75.
Zurück zum Zitat Jimeno A, Hallur G, Chan A, Zhang X, Cusatis G, Chan F, et al. Development of two novel benzoylphenylurea sulfur analogues and evidence that the microtubule-associated protein tau is predictive of their activity in pancreatic cancer. Mol Cancer Ther. 2007;6(5):1509–16.PubMedCrossRef Jimeno A, Hallur G, Chan A, Zhang X, Cusatis G, Chan F, et al. Development of two novel benzoylphenylurea sulfur analogues and evidence that the microtubule-associated protein tau is predictive of their activity in pancreatic cancer. Mol Cancer Ther. 2007;6(5):1509–16.PubMedCrossRef
76.
Zurück zum Zitat Lakhani NJ, Sarkar MA, Venitz J, Figg WD. 2-Methoxyestradiol, a promising anticancer agent. Pharmacotherapy. 2003;23(2):165–72.PubMedCrossRef Lakhani NJ, Sarkar MA, Venitz J, Figg WD. 2-Methoxyestradiol, a promising anticancer agent. Pharmacotherapy. 2003;23(2):165–72.PubMedCrossRef
77.
Zurück zum Zitat Chaturvedi P, George V, Shrestha N, Wang M, Dee MJ, Zhu X, et al. Immunotherapeutic HCW9218 augments anti-tumor activity of chemotherapy via NK cell-mediated reduction of therapy-induced senescent cells. Mol Ther. 2022;30(3):1171–87. Chaturvedi P, George V, Shrestha N, Wang M, Dee MJ, Zhu X, et al. Immunotherapeutic HCW9218 augments anti-tumor activity of chemotherapy via NK cell-mediated reduction of therapy-induced senescent cells. Mol Ther. 2022;30(3):1171–87.
78.
Zurück zum Zitat Nagaraju GP, Mezina A, Shaib WL, Landry J, El-Rayes BF. Targeting the Janus-activated kinase-2-STAT3 signalling pathway in pancreatic cancer using the HSP90 inhibitor ganetespib. Eur J Cancer. 2016;52:109–19.PubMedCrossRef Nagaraju GP, Mezina A, Shaib WL, Landry J, El-Rayes BF. Targeting the Janus-activated kinase-2-STAT3 signalling pathway in pancreatic cancer using the HSP90 inhibitor ganetespib. Eur J Cancer. 2016;52:109–19.PubMedCrossRef
79.
Zurück zum Zitat Ogunleye AZ, Piyawajanusorn C, Gonçalves A, Ghislat G, Ballester PJ. Interpretable machine learning models to predict the resistance of breast cancer patients to doxorubicin from their microRNA profiles. Adv Sci. 2022;9:2201501. Ogunleye AZ, Piyawajanusorn C, Gonçalves A, Ghislat G, Ballester PJ. Interpretable machine learning models to predict the resistance of breast cancer patients to doxorubicin from their microRNA profiles. Adv Sci. 2022;9:2201501.
Metadaten
Titel
NeRD: a multichannel neural network to predict cellular response of drugs by integrating multidimensional data
verfasst von
Xiaoxiao Cheng
Chong Dai
Yuqi Wen
Xiaoqi Wang
Xiaochen Bo
Song He
Shaoliang Peng
Publikationsdatum
01.12.2022
Verlag
BioMed Central
Erschienen in
BMC Medicine / Ausgabe 1/2022
Elektronische ISSN: 1741-7015
DOI
https://doi.org/10.1186/s12916-022-02549-0

Weitere Artikel der Ausgabe 1/2022

BMC Medicine 1/2022 Zur Ausgabe

Leitlinien kompakt für die Allgemeinmedizin

Mit medbee Pocketcards sicher entscheiden.

Seit 2022 gehört die medbee GmbH zum Springer Medizin Verlag

Facharzt-Training Allgemeinmedizin

Die ideale Vorbereitung zur anstehenden Prüfung mit den ersten 24 von 100 klinischen Fallbeispielen verschiedener Themenfelder

Mehr erfahren

Update Allgemeinmedizin

Bestellen Sie unseren Fach-Newsletter und bleiben Sie gut informiert.