Git Product home page Git Product logo

vasc's Introduction

<script src="https://polyfill.io/v3/polyfill.min.js?features=es6"></script> <script src="https://cdn.jsdelivr.net/npm/mathjax@3/es5/tex-mml-chtml.js"></script>

What is VASC? Exploring Dimension Reduction Built for RNAseq

Tiffany Hsieh & Tom Arnold


Below we will explore the VASC methodology, first by building an understanding of autoencoders, then variational autoencoders, and finally considering the novel VASC algorithm. The paper on which this is based is available here.

What is an Autoencoder?

In class, we discussed the feedforward neural network, and autoencoders are a special type of feedforward neural network. The autoencoder algorithm comprises two main elements: the encoder and the decoder.

image tooltip here

The encoder transforms the input data into a reduced dimensional representation, which is referred to as “latent space”; from the latent space, the decoder then reconstructs the output, which ideally should be as close as the input data.

Broadly, the loss function here is usually the reconstruction loss, which is the difference between the original input and the reconstructed output [1].

What is the Variational Autoencoder?

Before we jump into the topic of our paper, we first need to talk about the foundation of that algorithm, which is a specific type of autoencoder, the variational encoder (VAE).

Variational autoencoder was proposed in 2013 by Diederik P. Kingma and Max Welling at Google and Qualcomm [2]. The variational autoencoder is a special type of autoencoder that provides a probabilistic manner for describing an observation in latent space. Rather than a single encoding vector for the latent space, VAEs model two different vectors: a vector of means, “$$\mu$$,” and a vector of standard deviations, “$$\sigma$$”. This way, the VAEs allow us to interpolate and use random samples, which greatly expands their capabilities.

image tooltip here

Math behind VAE

Now that we understand the basic structure of VAE, let's dive deeper into the math [3]. We want to generate a distribution of the latent space $$z$$ from the observational data $$x$$, intuitively, we would do this: $$P(z|X) =\frac{P(X|z)P(z)}{P(X)}$$ However, the calculation of $$P(z|X)$$ is intractable. Therefore, we want to approximate $$P(z|X)$$ using the variational probability $$Q(z|X)$$ by minimizing the Kullback Leibler (KL) Divergence:

$$D[Q(z \mid X) | P(z \mid X)] = E_{z \sim Q} [\log Q(z \mid X) - \log P(z \mid X)]$$

By applying Bayes’ rule: $$\log P(X) - D[Q(z \mid X) | P(z \mid X)] = E_{z \sim Q} [\log P(X \mid z) - D[Q(z) | P(z)]]$$

where $$P(X)$$ is constant $$E_{z \sim Q}$$ represents the expectation over $$z$$ that is sampled from $$Q$$. Minimizing the KL divergence is equivalent to maximizing the right-hand part of the equation. The right-hand part has a natural autoencoder structure, with the encoder $$Q(z|X)$$ from $$X$$ to $$z$$ and the decoder $$P(X|z)$$ from $$z$$ to $$X$$.

Intuition of the paper

Before we learn the method of this paper, let's switch gears a bit. We need to know what this method is inspired by and what problem are the authors trying to solve.

RNA sequencing

RNA sequencing is a powerful tool for understanding the molecular mechanisms of cancer development and developing novel strategies for cancer prevention and treatment. There are generally two types of RNA sequencing techniques: bulk RNA sequencing and single-cell RNA sequencing (scRNA-seq) [4]. Bulk RNA sequencing measures the average gene expression across the population of various cells. With bulk RNA sequencing, you can compare the results of lung cancer patients with those of healthy ones [5]. However, the answers to that question may lie behind certain cell types.

Here is an analogy: bulk RNA sequencing is like a glass of smoothie; it has all kinds of fruits and vegetables, same as a blood sample, which is a mixture of different kinds of cell types, e.g., B cells, T cells. If we are particularly interested in the flavor characteristics of raspberry, it is difficult to do that with a glass of smoothie mixed with bananas, oranges, and pineapples. Therefore, this requires looking at the expression of genes in individual cells instead of an average representation. In that case, single-cell sequencing provides the potential to find molecular differences that are only linked to specific cell types.

As we learned from the analogy earlier, scRNA-seq provided a way to comprehensively characterize individual cells' transcriptional information, enabling us to gain full insight into the interplay of the transcripts.

However, scRNA-seq measurements typically suffer from large fractions of observed zeros, where a given gene in a given cell has no unique molecular identifiers or reads mapping to it. This is a very difficult challenge when it comes to dimension reduction. Therefore, this paper proposed a new method: deep variational autoencoder for scRNA-seq data (VASC) to tackle this issue.

Math & Code Behind VASC

Unlike VAE, which generally only has three steps, VASC can be broken down into 6 steps:

image tooltip here

Input layer

The input layer uses the expression matrix from scRNA-seq data. The input layer handles three major transformations: ensuring non-negativity, log-transformation, and re-scaling. These transformations are enabled by default given the algorithm's specific application for RNA-seq data for which these transformations are typically applied. The log-transformation and scaling help compress the dynamic range of the input data which can help with optimization in a non-convex space such as a neural network.

expr[expr<0] = 0.0
if log:
    expr = np.log2( expr + 1 )
if scale:
    for i in range(expr.shape[0]):
        expr[i,:] = expr[i,:] / np.max(expr[i,:])

Dropout layer

A dropout layer is added immediately after the input layer, with the dropout rate set as 0.5. This layer randomly sets some features to zeros during the encoding phase, to increase the performance in model learning. This layer forces subsequent layers to learn to avoid dropout noises. Mathematically, if we denote the input to the dropout layer as $$ h_0 $$, and the output as $$ h_{\text{drop}0} $$, then: $$ h{\text{drop}_0}[i,j] = h_0[i,j] \times m[i,j] $$ where $$ m[i,j] \sim \text{Bernoulli}(0.5) $$ for each $$ i, j $$. In other words, each element of $$ h_0 $$ is independently set to 0 with 50% probability [6].

h0 = Dropout(0.5)(expr_in)

Encoder network

The encoder network is designed as a three-layer fully-connected neural network with decreasing dimensions 512, 128, and 32. L1-norm regularization was added for the weights in this layer, which penalizes the sparsity of the model. The next two layers are accompanied by ReLU activation, which allows the model to learn complex non-linear relationships. As noted above, the encoder produces the learned latent representation of the original data.

h1 = Dense( units=512,name='encoder_1',kernel_regularizer=regularizers.l1(0.01) )(h0)
h2 = Dense( units=128,name='encoder_2' )(h1)  
h3 = Dense( units=32,name='encoder_3' )(h2_relu)

Latent Sampling

The latent sampling layer generates samples from the learned latent representation. This layer takes the output from the encoder network and uses it to parameterize a probability distribution in the latent space, from which samples are then drawn. The latent variables $$ z $$ are modeled using a multivariate Gaussian distribution. The encoder network outputs the parameters of this distribution: the mean vector $$ \mu $$ and the covariance matrix $$ \Sigma $$ or, if the var flag is enabled, the log-variances which are exponentiated to get the diagonal of $$ \Sigma$$ .

Mathematically, the latent distribution is $$ z \sim \mathcal{N}(\mu, \Sigma) $$ where $$ \mu = \mu(X) $$ and $$ \Sigma = \Sigma(X) $$ are functions of the input data X, learned by the encoder network. In the code, the latent sampling layer is implemented as follows.

z_mean = Dense( units= self.latent ,name='z_mean' )(h3_relu)

The line above uses a fully-connected layer to map the output of the encoder (h3_relu) to the mean vector $$ \mu $$ of the latent Gaussian distribution.

if self.var:
    z_log_var = Dense( units=2,name='z_log_var' )(h3_relu)
    z_log_var = Activation( 'softplus' )(z_log_var)
    z = Lambda(sampling, output_shape=(self.latent,))([z_mean,z_log_var])
else:
    z = Lambda(sampling, output_shape=(self.latent,))([z_mean])

The code above handles the sampling process. If the var parameter is True, another fully-connected layer is used to output the log-variances of the Gaussian, which are then passed through a softplus activation to ensure they are positive (since variances must be non-negative). The sampling function is then called with both $$ \mu $$ and $$ \Sigma $$. (If var is False, the sampling function is called with only $$ \mu $$, and a fixed unit variance is used.)

def sampling(args):
    epsilon_std = 1.0

    args = tf.convert_to_tensor(args)

    if args.shape[0] == 2:
        z_mean, z_log_var = args
        epsilon = K.random_normal(shape=K.shape(z_mean),
                              mean=0.,
                              stddev=epsilon_std)
    #
        return z_mean + K.exp( z_log_var / 2 ) * epsilon
    else:
        z_mean = args[0]
        epsilon = K.random_normal(shape=K.shape(z_mean),
                              mean=0.,
                              stddev=epsilon_std)
        return z_mean + K.exp( 1.0 / 2 ) * epsilon

The function above implements a sampling reparameterization trick, which allows for more efficient gradient computation during training. Instead of directly sampling from $$ \mathcal{N}(\mu, \Sigma) $$, it samples a standard Gaussian noise $$\varepsilon \sim \mathcal{N}(0, 1) $$ and then computes $$ z = \mu + \Sigma^{1/2} \times \varepsilon $$. This is equivalent to sampling from $$ \mathcal{N}(\mu, \Sigma) $$ but allows for backpropagation through the sampling step.

Decoder network

The decoder network maps the latent representation $$z$$ back to the original gene expression space, thereby reconstructing the input data. It is a mirror image of the encoder network. It consists of a three-layer fully-connected neural network with dimensions of hidden units 32, 128, and 512, respectively, and an output layer. The decoder transforms the ​​latent representation into a higher-dimensional output that matches the dimensions of the input gene expression profile.

decoder_h1 = Dense( units=32,name='decoder_1' )(z)
decoder_h2 = Dense( units=128,name='decoder_2' )(decoder_h1_relu)
decoder_h3 = Dense( units=512,name='decoder_3' )(decoder_h2_relu)
expr_x = Dense(units=self.in_dim,activation='sigmoid')(decoder_h3_relu)

The final layer of the decoder maps the output of decoder_h3 to the original dimension of the gene expression profile.

Mathematically, if we call the function represented by the decoder network as $$ g(\cdot) $$, then $$ \hat{X} = g(z) $$ when $$ \hat{X} $$ is the reconstructed gene expression profile. The goal of the decoder is to make $$ \hat{X} $$ as close as possible to the original input $$ X $$. This is achieved through the optimization of the reconstruction term in the loss function (a mix of KL divergence and binary cross-entropy between $$ X $$ and $$ \hat{X} $$).

ZI layer

An additional ZI layer is added after the decoder network, and could be considered the VASC ‘secret sauce.’ The ZI layer models the dropout events by setting some decoded expression values to zero based on a double exponential distribution. The dropout events are modeled by the probability $$p_{ij} = e^{-\tilde{y}^2} $$.

def sampling_gumbel(shape,eps=1e-8):
    u = K.random_uniform( shape )
    return -K.log( -K.log(u+eps)+eps )

def compute_softmax(logits,temp):
    z = logits + sampling_gumbel( K.shape(logits) )
    return K.softmax( z / temp )

def gumbel_softmax(args):
    logits,temp = args
    return compute_softmax(logits,temp)

Since back-propagation cannot deal with stochastic units, new approach was proposed. A Gumbel- softmax distribution, $$s = \frac{\exp\left(\frac{\log p + g_0}{\tau}\right)}{\exp\left(\frac{\log p + g_0}{\tau}\right) + \exp\left(\frac{\log q + g_1}{\tau}\right)}$$ was introduced to overcome this issue, where $$p $$ is the probability for dropout and $$q=1-p$$,$$g_0$$,$$g_1$$ were sampled from a $$Gumbel \sim (0,1)$$ distribution. The samples could then be obtained by first drawing an auxiliary sample $$u \sim Uniform (0,1)$$ and then computing $$g=-log(-log u)$$

expr_x_drop = Lambda(lambda x: -x ** 2)(expr_x)
expr_x_drop_p = Lambda( lambda x:K.exp(x) )(expr_x_drop)
expr_x_nondrop_p = Lambda( lambda x:1-x )( expr_x_drop_p )
logits = merge( [expr_x_drop_log,expr_x_nondrop_log],mode='concat',concat_axis=-1 )
samples = Lambda( gumbel_softmax,output_shape=(self.in_dim,2,) )( [logits,temp_] )   

In the code above we see expr_x is the output of the decoder network (without the sigmoid activation). expr_x_drop computes $$-y_{ij}^2$$, and expr_x_drop_p computes the dropout probability $$p_{ij}$$. expr_x_nondrop_p computes the probability of not being a dropout $$(1 - p_{ij})$$. The code then computes the log-probabilities and concatenates them. These log-probabilities are used to compute the final output of the ZI layer using the Gumbel-Softmax trick. Said another way, the Gumbel-Softmax trick is used to generate a differentiable approximation to sampling from a discrete distribution, enabling gradient backpropagation through the ZI layer.

In sum, the ZI layer models the dropout events in the input RNA-seq data by estimating a dropout probability for each gene based on the latent representation $$z$$. It then uses the Gumbel-Softmax trick to sample binary “dropout masks” that are multiplied element-wise with the decoder output.

Loss function

There are two parts of the loss function. The first part was computed by binary cross-entropy loss function, since the scale of the data is [0,1], which can be seen to be the ‘reconstruction loss.’ Binary cross-entropy can be interpreted as the probability of each gene being expressed.

The second part is the same as minimizing the Kullback–Leibler (KL) divergence typically seen in VAE. The reconstruction loss ensures that the autoencoder is able to effectively reconstruct the input data from the latent representation, while the KL divergence loss acts as a regularizer, encouraging the learned latent distribution to be close to a prior distribution (assumed to be a standard Gaussian).

def vae_loss(self, x, x_decoded_mean):
    xent_loss = in_dim * metrics.binary_crossentropy(x, x_decoded_mean)
    if var_:
        kl_loss = - 0.5 * K.sum(1 + z_log_var - K.square(z_mean) - K.exp(z_log_var), axis=-1)
    else:
        kl_loss = - 0.5 * K.sum(1 + 1 - K.square(z_mean) - K.exp(1.0), axis=-1)
    return K.mean(xent_loss + kl_loss)

Mathematically, the total loss for a single sample can be written as:

$$ L = - E_{q(z|x)}(\log p(x|z)) + \text{KL}(q(z|x) \parallel p(z)) $$

When $$ q(z \mid x) $$ is the learned latent distribution (the encoder output), $$ p(x \mid z) $$ is the decoder output distribution, and $$ p(z) $$ is the prior distribution on the latent space. The first term is the negative expected log-likelihood of the data under the decoder output distribution, which corresponds to the reconstruction loss. The second term is the KL divergence between the learned latent distribution and the prior, which acts as a regularizer.

Optimization

The entire structure as delineated above is optimized by a variant of the stochastic gradient descent optimization algorithm that we learned in class known as ​​RMSprop. RMSprop has an adaptive learning rate (analogous to momentum) for each parameter similar to Adam, but without the frictional component. It should also be noted that the algorithm uses batch processing to avoid overfitting and promote faster learning.

opt = RMSprop( lr=0.0001 )
vae.compile( optimizer=opt,loss=None )

This process is repeated for a fixed number of iterations, or until a stopping criterion is met. The training stops if there is no obvious (greater than 0.1) decrease in the loss function within a given number of epochs. In terms of code, the training loop is set to run for a maximum of 'epoch' iterations, but will stop early if the loss does not improve significantly for 'patience' epochs in a row.

        if e % patience == 1:
            print("Epoch %d/%d" % (e, epochs))
            print("Loss:" + str(train_loss))
            print("current loss: ", cur_loss, "previous loss:", prev_loss)
            if abs(cur_loss - prev_loss) < 0.1:
                print('current loss - prev loss < 0.1, breaking')
                break
            prev_loss = train_loss

Experimental Results from Our Implementation

To further explore, understand the algorithm, and verify results, we implemented our own version of the VASC algorithm. This algorithm is functionally identical to the original, with updates only for modernization with respect to the libraries used. We employ the same procedure as in the original paper, where a dimension reduction technique is applied, tSNE is used for visualization, and finally k-Means is used to predict cell type assignments.

Visualization of the Original Data (Using tSNE)

image tooltip here{:width=350}

For future comparison, we produce a two dimensional visualization of the raw input data (test data provided by the authors, the Biase dataset). The example dataset contains 56 cells and 25,733 genes. We note the clear separation between cell types.

VASC Output: Latent Space (Visualization & Metrics)

image tooltip here{:width=350}

Normalized Mutual Information (NMI): 0.9232425411905947
Adjusted Rand Index (ARI): 0.9217481265281982
Homogeneity: 0.9071668964991539
Completeness: 0.9398982080777898

First we note that we are able to produce the exact values cited by the original authors in Figure 3, confirming our correct implementation of the algorithm. By examining the latent representation of the data, we are exploring how VASC is mapping the scRNA-seq data into a the 2-dimensional latent space. It is remarkably able to retain more than 92% of the information from over 25,000 genes in just two latent dimensions. However, we do note that as with all neural networks these results are somewhat volatile with respect to the local optima in the non-convex space as well as highly sensitive to hyperparameters. Extensive experimentation was required to produce this result. Lastly, VASC was the slowest of the three tested algorithms, by a wide margin.

tSNE Output: Reduced Data (Visualization & Metrics)

image tooltip here{:width=350}

Normalized Mutual Information (NMI): 0.724993024280294
Adjusted Rand Index (ARI): 0.6603062495667171
Homogeneity: 0.7323475030067922
Completeness: 0.717784789199681

Here we note that tSNE does not preform as well as VASC or PCA below. However, it performs much better than as noted in the original paper, indicating a potential mistake on the part of the authors. We might expect tSNE to not preform well on highly sparse data given that stochastic neighbor embedding relies on local relationships to identify patters in the data.

PCA Output: Reduced Data (Visualization & Metrics)

image tooltip here{:width=350}

Normalized Mutual Information (NMI): 0.8628264996152113
Adjusted Rand Index (ARI): 0.8697708506572868
Homogeneity: 0.8419827834744833
Completeness: 0.8847284055799015

Finally, we see that PCA preforms well, but still not as well as VASC. This PCA is using the first 50 principal components, which makes it all the more remarkable that VASC was able to outperform it with only two latent dimensions.

Lastly we note that the above comparisons are all with respect to the dimensionally reduced space. That is to say, none of the representations attempt to project the reduced data back to the original space to quantify the amount of information lost.

As with PCA, the reconstructed data VASC represents the information lost in dimension reduction. In PCA, reconstructed data is assessing the loss of information, representing how well the given principal components can approximate the original data after reducing its dimensionality. This process is linear and aims to preserve the maximum variance. For a VAE like VASC, reconstructed data serves to evaluate the model’s ability to generate or simulate the original data from a probabilistically modeled latent space. This reconstruction assesses how well the VAE has learned the underlying distribution of the data with their non-linear encoding-decoding process, but is also about assessing the loss of information.

Therefore while the authors fairly compare the latent space representing of VASC with the reduced dimension space of other algorithms, it may be more fair to compare reconstructed data in all contexts. Below we proceed with an example.

VASC Output: Reconstructed Data (Visualization & Metrics)

image tooltip here{:width=350}

Normalized Mutual Information (NMI): 0.8824205399613051
Adjusted Rand Index (ARI): 0.901202031763618
Homogeneity: 0.8697210308789426
Completeness: 0.895496416560308

PCA Output: Reconstructed Data (Visualization & Metrics)

image tooltip here{:width=350}

Normalized Mutual Information (NMI): 0.8628264996152113
Adjusted Rand Index (ARI): 0.8697708506572868
Homogeneity: 0.8419827834744833
Completeness: 0.8847284055799015

We see that VASC is still superior to PCA, but the margin has shrunk. We note that PCA is a linear transformation that reduces dimensionality by projecting the data onto the principal components. When we reduce and then reconstruct the data (using a subset of principal components), the reconstruction involves projecting back onto the original space using the same set of principal components. This projection back and forth theoretically incurs a loss of information only along the least significant components (those not retained).

However, VASC loses some information in the decoder phase as a consequence of its architecture discussed above. For this reason, it would have been better for the authors to compare reconstructed data, not merely reduced data. Nevertheless, VASC is still able to represent extremely high-dimension data in just two latent dimensions; a huge accomplishment despite the drawbacks.

Experimental Results from Paper

The authors tested the visualization performance of VASC together with four state-of-the-art dimension reduction methods: PCA, ZIFA, t-SNE, and SIMLR. They used 20 datasets with different number of cells included and sequencing protocols used. Performance assessments were measured using k-means clustering.

image tooltip here

The top panel (Figure 3A) shows the NMI and ARI values for each method on each dataset. NMI stands for normalized mutual information, which is calculated as $$NMI(P, T) = \frac{MI(P, T)}{\sqrt{H(P) H(T)}}$$, where $$P$$ is the predicted clustering results, and $$T$$ is the known cell types, and $$H(P)$$ and $$H(T)$$ are the entropy of $$P$$ and $$T$$, the mutual information between them as $$MI(P,T)$$.

ARI stands for adjusted rand index, it is calculated as

$$\text{ARI} = \frac{\sum_{ij} \binom{n_{ij}}{2} - \left[\sum_i \binom{a_i}{2} \sum_j \binom{b_j}{2}\right] \Big/ \binom{n}{2}} {\frac{1}{2} \left[\sum_i \binom{a_i}{2} + \sum_j \binom{b_j}{2}\right] - \left[\sum_i \binom{a_i}{2} \sum_j \binom{b_j}{2}\right] \Big/ \binom{n}{2}}$$

where $$n$$ is the total number of samples, $$a_i$$ is the number of samples appearing in the $$i$$-th cluster of $$P$$, $$b_j$$ is the number of samples appearing in the $$j$$-th types of $$T$$, and $$n_{ij}$$ is the number of overlaps between the $$i$$-th cluster of $$P$$ and the $$j$$-th type and $$T$$.

From the results from the top panel, we see that VASC outperformed the other methods in terms of NMI and ARI in most cases (best performances achieved on 15 and 17 out of the 20 datasets, respectively).

The lower panel shows the statistics of the ranks of the compared methods based on NMI and ARI values. For each dataset, NMI and ARI values given by different algorithms were ranked in descending order, with rank 1 indicative of the highest NMI or ARI values. The number of ranks achieved by these algorithms in the 20 datasets is then counted for distribution. As we can see from the results, VASC always ranked in the top two methods of all the tested datasets in terms of NMI and ARI.

Conclusions

Overall, the results from the paper suggested that VASC has broad compatibility with various kinds of scRNA-seq datasets and performs better than PCA and ZIFA, especially when the sample sizes are larger. VASC achieves superior performance in most cases and is broadly suitable for different datasets with different data structures in the original space. We were able to verify these results independently, and in our independent analysis we were able to determine that while VASC results in superior representations of the data in many cases, it is highly dependent on parameter tuning and is computationally expensive (especially relative to more straightforward linear operations like PCA).

References

[1] Deep Learning (Ian J. Goodfellow, Yoshua Bengio and Aaron Courville), MIT Press, 2016.

[2] Kingma, D. P., & Welling, M. (2019). An Introduction to Variational Autoencoders. ArXiv. https://doi.org/10.1561/2200000056

[3] Wang, Dongfang, and Jin Gu. "VASC: dimension reduction and visualization of single-cell RNA-seq data by deep variational autoencoder." Genomics, Proteomics and Bioinformatics 16.5 (2018): 320-331.

[4] Li, X., Wang, CY. From bulk, single-cell to spatial RNA sequencing. Int J Oral Sci 13, 36 (2021). https://doi.org/10.1038/s41368-021-00146-0

[5] Yu X, Abbas-Aghababazadeh F, Chen YA, Fridley BL. Statistical and Bioinformatics Analysis of Data from Bulk and Single-Cell RNA Sequencing Experiments. Methods Mol Biol. 2021;2194:143-175. doi: 10.1007/978-1-0716-0849-4_9. PMID: 32926366; PMCID: PMC7771369.

[6] Srivastava, Nitish, et al. "Dropout: A Simple Way to Prevent Neural Networks from Overfitting." Journal of Machine Learning Research, vol. 15, no. 56, 2014, pp. 1929-1958, http://jmlr.org/papers/v15/srivastava14a.html.

Appendix: Code

VASC code can be found at this link.


Thank you for reading!

vasc's People

Contributors

tomrannosaurus avatar

Watchers

 avatar

Recommend Projects

  • React photo React

    A declarative, efficient, and flexible JavaScript library for building user interfaces.

  • Vue.js photo Vue.js

    🖖 Vue.js is a progressive, incrementally-adoptable JavaScript framework for building UI on the web.

  • Typescript photo Typescript

    TypeScript is a superset of JavaScript that compiles to clean JavaScript output.

  • TensorFlow photo TensorFlow

    An Open Source Machine Learning Framework for Everyone

  • Django photo Django

    The Web framework for perfectionists with deadlines.

  • D3 photo D3

    Bring data to life with SVG, Canvas and HTML. 📊📈🎉

Recommend Topics

  • javascript

    JavaScript (JS) is a lightweight interpreted programming language with first-class functions.

  • web

    Some thing interesting about web. New door for the world.

  • server

    A server is a program made to process requests and deliver data to clients.

  • Machine learning

    Machine learning is a way of modeling and interpreting data that allows a piece of software to respond intelligently.

  • Game

    Some thing interesting about game, make everyone happy.

Recommend Org

  • Facebook photo Facebook

    We are working to build community through open source technology. NB: members must have two-factor auth.

  • Microsoft photo Microsoft

    Open source projects and samples from Microsoft.

  • Google photo Google

    Google ❤️ Open Source for everyone.

  • D3 photo D3

    Data-Driven Documents codes.