Spatial functional analysis

Spatial transcriptomics technologies yield many molecular readouts that are hard to interpret by themselves. One way of summarizing this information is by inferring biological activities from prior knowledge.

In this notebook we showcase how to use decoupler for transcription factor and pathway activity inference from a human data-set. The data consists of a 10X Genomics Visium slide of a human lymph node and it is available at their website.

Note

This tutorial assumes that you already know the basics of decoupler. Else, check out the Usage tutorial first.

Loading packages

First, we need to load the relevant packages, scanpy to handle RNA-seq data and decoupler to use statistical methods.

[1]:
import scanpy as sc
import decoupler as dc

# Only needed for processing
import numpy as np
import pandas as pd

# Plotting options, change to your liking
sc.settings.set_figure_params(dpi=200, frameon=False)
sc.set_figure_params(dpi=200)
sc.set_figure_params(figsize=(4, 4))

Loading the data

We can download the data easily using scanpy:

[2]:
adata = sc.datasets.visium_sge(sample_id="V1_Human_Lymph_Node")
adata.var_names_make_unique()
adata
[2]:
AnnData object with n_obs × n_vars = 4035 × 36601
    obs: 'in_tissue', 'array_row', 'array_col'
    var: 'gene_ids', 'feature_types', 'genome'
    uns: 'spatial'
    obsm: 'spatial'

QC, projection and clustering

Here we follow the standard pre-processing steps as described in the scanpy vignette. These steps carry out the selection and filtration of spots based on quality control metrics and data normalization.

Note

This is just an example, these steps should change depending on the data.

[3]:
# Basic filtering
sc.pp.filter_cells(adata, min_genes=200)
sc.pp.filter_genes(adata, min_cells=10)

# Normalize the data
sc.pp.normalize_total(adata, target_sum=1e4)
sc.pp.log1p(adata)

Then we group spots based on the similarity of their transcription profiles. We can visualize the obtained clusters on a UMAP or directly at the tissue:

[4]:
# Store norm counts
adata.layers['log_norm'] = adata.X.copy()

# Identify highly variable genes
sc.pp.highly_variable_genes(adata)

# Scale the data
sc.pp.scale(adata, max_value=10)

# Generate PCA features
sc.tl.pca(adata, svd_solver='arpack')

# Restore X to be norm counts
dc.swap_layer(adata, 'log_norm', X_layer_key=None, inplace=True)

# Compute distances in the PCA space, and find spot neighbors
sc.pp.neighbors(adata)

# Run leiden clustering algorithm
sc.tl.leiden(adata)
adata.obs['leiden'] = ['Clust. {0}'.format(i) for i in adata.obs['leiden']]
[5]:
sc.pl.spatial(adata, color=[None, 'leiden'], size=1.5, wspace=0, frameon=False)
../_images/notebooks_spatial_10_0.png

Spatial Connectivity

Additionally, we can incorporate spatial information in our analyses by accounting for the spatial connectivity of our observations (in this case spots). By weighting the spatial proximity of our spots to the neighboring ones, we can transform our data to incorporate spatial context.

Note

This section is optional, it is also valid to compute enrichment scores without taking into account the spatial component.

To generate spatial weights based on spatial proximity we will use liana. More information is available in this vignette.

[6]:
import liana as li

li.ut.spatial_neighbors(
    adata,
    bandwidth=150,
    cutoff=0.1,
    kernel='gaussian',
    set_diag=True,
    standardize=True
)
[7]:
# Plot the spatial weights of one spot in our object
adata.obs['conn'] = adata.obsp['spatial_connectivities'][0].A.ravel()
sc.pl.spatial(adata, color='conn', size=1.5, frameon=False)
../_images/notebooks_spatial_13_0.png

We can see that for a given spot, the spatial weight is at its maximum on itself and decreases the more distant other spots are. Now we can spatially weight the gene expression values:

[8]:
# Update X with spatially weighted gene exression
adata.X = adata.obsp['spatial_connectivities'].A.dot(adata.X.A)

We can plot the original log-normalized counts with the new spatialy transformed values:

[9]:
genes = ['IFIT3', 'IGKC']

# Log-normalized counts
sc.pl.spatial(adata, color=genes, size=1.5, frameon=False, layer='log_norm')

# Spatially weighted gene expression
sc.pl.spatial(adata, color=genes, size=1.5, frameon=False)
../_images/notebooks_spatial_17_0.png
../_images/notebooks_spatial_17_1.png

Transcription factor activity inference

The first functional analysis we can perform is to infer transcription factor (TF) activities from our transcriptomics data. We will need a gene regulatory network (GRN) and a statistical method.

CollecTRI network

CollecTRI is a comprehensive resource containing a curated collection of TFs and their transcriptional targets compiled from 12 different resources. This collection provides an increased coverage of transcription factors and a superior performance in identifying perturbed TFs compared to our previous DoRothEA network and other literature based GRNs. Similar to DoRothEA, interactions are weighted by their mode of regulation (activation or inhibition).

For this example we will use the human version (mouse and rat are also available). We can use decoupler to retrieve it from omnipath. The argument split_complexes keeps complexes or splits them into subunits, by default we recommend to keep complexes together.

Note

In this tutorial we use the network CollecTRI, but we could use any other GRN coming from an inference method such as CellOracle, pySCENIC or SCENIC+.

[10]:
net = dc.get_collectri(organism='human', split_complexes=False)
net
Downloading data from `https://omnipathdb.org/queries/enzsub?format=json`
Downloading data from `https://omnipathdb.org/queries/interactions?format=json`
Downloading data from `https://omnipathdb.org/queries/complexes?format=json`
Downloading data from `https://omnipathdb.org/queries/annotations?format=json`
Downloading data from `https://omnipathdb.org/queries/intercell?format=json`
Downloading data from `https://omnipathdb.org/about?format=text`
[10]:
source target weight PMID
0 MYC TERT 1 10022128;10491298;10606235;10637317;10723141;1...
1 SPI1 BGLAP 1 10022617
2 SMAD3 JUN 1 10022869;12374795
3 SMAD4 JUN 1 10022869;12374795
4 STAT5A IL2 1 10022878;11435608;17182565;17911616;22854263;2...
... ... ... ... ...
43173 NFKB hsa-miR-143-3p 1 19472311
43174 AP1 hsa-miR-206 1 19721712
43175 NFKB hsa-miR-21-5p 1 20813833;22387281
43176 NFKB hsa-miR-224-5p 1 23474441;23988648
43177 AP1 hsa-miR-144 1 23546882

43178 rows × 4 columns

Activity inference with Univariate Linear Model (ULM)

To infer TF enrichment scores we will run the Univariate Linear Model (ulm) method. For each spot in our slide (adata) and each TF in our network (net), it fits a linear model that predicts the observed gene expression based solely on the TF’s TF-Gene interaction weights. Once fitted, the obtained t-value of the slope is the score. If it is positive, we interpret that the TF is active and if it is negative we interpret that it is inactive.

7dfd103048cd4c5195741cec4a6479ae

To run decoupler methods, we need an input matrix (mat), an input prior knowledge network/resource (net), and the name of the columns of net that we want to use.

[11]:
dc.run_ulm(
    mat=adata,
    net=net,
    source='source',
    target='target',
    weight='weight',
    verbose=True,
    use_raw=False
)
Running ulm on mat with 4032 samples and 19813 targets for 714 sources.

The obtained scores (ulm_estimate) and p-values (ulm_pvals) are stored in the .obsm key:

[12]:
adata.obsm['ulm_estimate']
[12]:
ABL1 AHR AIP AIRE AP1 APEX1 AR ARID1A ARID1B ARID3A ... ZNF382 ZNF384 ZNF395 ZNF423 ZNF436 ZNF699 ZNF76 ZNF804A ZNF91 ZXDC
AAACAAGTATCTCCCA-1 2.506540 2.993418 -0.224612 1.088716 8.932042 3.467653 8.088826 0.097465 1.640161 1.590863 ... -2.141593 0.041473 0.137326 -1.132138 1.519211 2.201835 0.106273 -0.554477 1.577712 5.126571
AAACAATCTACTAGCA-1 1.861013 2.003174 -0.234753 0.840002 7.381738 2.214163 7.312593 -0.160331 1.457745 0.701439 ... -1.719987 -0.552122 -0.236105 -1.066291 1.508229 1.620679 0.599542 -0.930530 1.557906 4.704382
AAACACCAATAACTGC-1 2.365444 3.012575 -0.247704 1.302476 9.036439 3.478431 8.182927 0.190212 1.234260 1.334810 ... -1.935273 -0.352290 0.110648 -1.090288 1.026727 2.036324 0.634918 -0.373306 2.066460 4.922176
AAACAGAGCGACTCCT-1 1.333631 3.089248 -0.161767 1.122681 8.437885 3.095499 7.818810 -0.074081 2.859876 2.275510 ... -1.949705 -0.472413 1.141847 -1.144544 1.538650 1.718085 1.172972 0.167259 1.726077 4.207001
AAACAGCTTTCAGAAG-1 1.444055 2.606838 -0.424777 0.746851 7.309249 3.678788 6.895603 -0.038943 2.916240 2.473541 ... -1.677928 -0.270460 -0.091756 -1.113962 0.842817 1.549942 0.918052 0.020900 1.605347 4.549479
... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ... ...
TTGTTTCACATCCAGG-1 2.178375 2.633951 -0.171126 0.774759 9.022694 3.937260 7.809430 0.121977 1.853033 1.935024 ... -1.907115 -0.074763 -0.093024 -1.062878 1.136265 1.897447 0.283017 -0.189876 1.996677 5.542336
TTGTTTCATTAGTCTA-1 2.148535 3.249435 -0.164014 1.266401 9.157030 3.788883 8.210668 -0.031493 1.914353 1.334394 ... -1.997858 -0.149779 -0.424333 -1.120992 1.545704 2.204408 0.580634 -0.393936 2.131220 5.359280
TTGTTTCCATACAACT-1 2.471877 3.029328 0.003349 1.515853 9.260170 3.733936 8.150513 0.272917 2.101049 1.329936 ... -2.033781 -0.252704 -0.093840 -1.051247 1.592259 1.741617 0.421107 -0.850126 1.839918 5.194634
TTGTTTGTATTACACG-1 2.095661 2.986843 -0.117071 1.187332 9.163303 3.595011 7.863039 0.219370 1.258791 1.072886 ... -2.060137 -0.159512 0.151495 -1.038933 1.136436 1.912966 0.376731 -0.870485 2.227823 5.392848
TTGTTTGTGTAAATTC-1 1.798320 2.622384 -0.314093 1.126644 8.291216 2.921452 7.545118 -0.136264 1.678849 1.517110 ... -1.643596 -0.724133 -0.226429 -1.170998 1.248425 1.649105 0.851030 -0.691052 1.609469 4.503331

4032 rows × 714 columns

Note: Each run of run_ulm overwrites what is inside of ulm_estimate and ulm_pvals. if you want to run ulm with other resources and still keep the activities inside the same AnnData object, you can store the results in any other key in .obsm with different names, for example:

[13]:
adata.obsm['collectri_ulm_estimate'] = adata.obsm['ulm_estimate'].copy()
adata.obsm['collectri_ulm_pvals'] = adata.obsm['ulm_pvals'].copy()
adata
[13]:
AnnData object with n_obs × n_vars = 4032 × 19813
    obs: 'in_tissue', 'array_row', 'array_col', 'n_genes', 'leiden', 'conn'
    var: 'gene_ids', 'feature_types', 'genome', 'n_cells', 'highly_variable', 'means', 'dispersions', 'dispersions_norm', 'mean', 'std'
    uns: 'spatial', 'log1p', 'hvg', 'pca', 'neighbors', 'leiden', 'leiden_colors'
    obsm: 'spatial', 'X_pca', 'ulm_estimate', 'ulm_pvals', 'collectri_ulm_estimate', 'collectri_ulm_pvals'
    varm: 'PCs'
    layers: 'log_norm'
    obsp: 'distances', 'connectivities', 'spatial_connectivities'

Visualization

To visualize the obtained scores, we can re-use many of scanpy’s plotting functions. First though, we need to extract them from the adata object.

[14]:
acts = dc.get_acts(adata, obsm_key='collectri_ulm_estimate')
acts
[14]:
AnnData object with n_obs × n_vars = 4032 × 714
    obs: 'in_tissue', 'array_row', 'array_col', 'n_genes', 'leiden', 'conn'
    uns: 'spatial', 'log1p', 'hvg', 'pca', 'neighbors', 'leiden', 'leiden_colors'
    obsm: 'spatial', 'X_pca', 'ulm_estimate', 'ulm_pvals', 'collectri_ulm_estimate', 'collectri_ulm_pvals'

dc.get_acts returns a new AnnData object which holds the obtained activities in its .X attribute, allowing us to re-use many scanpy functions, for example:

[15]:
sc.pl.spatial(
    acts,
    color=['IRF1', 'leiden'],
    cmap='RdBu_r',
    size=1.5,
    vcenter=0,
    frameon=False
)
sc.pl.violin(
    acts,
    keys='IRF1',
    groupby='leiden',
    rotation=90
)
../_images/notebooks_spatial_29_0.png
../_images/notebooks_spatial_29_1.png

Here we observe the activity infered for IRF1 across spots. Interestingly, IRF1 is a known TF crucial for interferons-mediated signaling pathways. The inference of activities from “foot-prints” of target genes is more informative than just looking at the molecular readouts of a given TF, as an example here is the gene expression of IRF1, which is not very informative by itself:

[16]:
sc.pl.spatial(
    adata,
    color=['IRF1', 'leiden'],
    size=1.5,
    frameon=False
)
sc.pl.violin(
    adata,
    keys='IRF1',
    groupby='leiden',
    rotation=90
)
../_images/notebooks_spatial_31_0.png
../_images/notebooks_spatial_31_1.png

Exploration

Let’s identify which are the top TF per cluster. We can do it by using the function dc.rank_sources_groups, which identifies marker TFs using the same statistical tests available in scanpy’s scanpy.tl.rank_genes_groups.

[17]:
df = dc.rank_sources_groups(acts, groupby='leiden', reference='rest', method='t-test_overestim_var')
df
[17]:
group reference names statistic meanchange pvals pvals_adj
0 Clust. 0 rest GSC 26.693179 0.720106 3.805158e-127 1.358442e-124
1 Clust. 0 rest ATF6 26.564711 0.806207 3.065348e-126 7.295529e-124
2 Clust. 0 rest XBP1 26.303573 0.776220 2.330251e-124 4.159497e-122
3 Clust. 0 rest TBXT 24.935113 0.716476 5.946454e-114 8.491537e-112
4 Clust. 0 rest PRDM1 23.248887 0.567659 8.897234e-101 1.058771e-98
... ... ... ... ... ... ... ...
7135 Clust. 9 rest APEX1 -12.681607 -0.542233 7.593762e-28 2.464521e-26
7136 Clust. 9 rest RFX1 -12.708333 -0.876288 3.256949e-28 1.162731e-26
7137 Clust. 9 rest PREB -12.731301 -0.518973 3.112465e-28 1.162731e-26
7138 Clust. 9 rest PITX1 -14.683193 -0.452296 1.235997e-34 5.515636e-33
7139 Clust. 9 rest KAT2B -20.452434 -0.554208 7.525725e-51 7.676239e-49

7140 rows × 7 columns

We can then extract the top 3 markers per cluster:

[18]:
n_markers = 3
source_markers = df.groupby('group').head(n_markers).groupby('group')['names'].apply(lambda x: list(x)).to_dict()
source_markers
[18]:
{'Clust. 0': ['GSC', 'ATF6', 'XBP1'],
 'Clust. 1': ['RFX5', 'NFATC3', 'SPIB'],
 'Clust. 2': ['STAT4', 'TRERF1', 'FOXP1'],
 'Clust. 3': ['NFKB2', 'JDP2', 'PBX1'],
 'Clust. 4': ['THRA', 'TTF1', 'HMGA2'],
 'Clust. 5': ['ARID3A', 'ARID1B', 'E2F4'],
 'Clust. 6': ['SPI1', 'CEBPG', 'CREB1'],
 'Clust. 7': ['MECOM', 'NFIX', 'ERG'],
 'Clust. 8': ['E2F3', 'E2F2', 'E2F1'],
 'Clust. 9': ['POU2F3', 'ZNF148', 'IRF9']}

We can plot the obtained markers:

[19]:
sc.pl.matrixplot(acts, source_markers, 'leiden', dendrogram=True, standard_scale='var',
                 colorbar_title='Z-scaled scores', cmap='RdBu_r')
WARNING: dendrogram data not found (using key=dendrogram_leiden). Running `sc.tl.dendrogram` with default parameters. For fine tuning it is recommended to run `sc.tl.dendrogram` independently.
../_images/notebooks_spatial_37_1.png

We can visualize again on the tissue some marker TFs:

[20]:
sc.pl.spatial(
    acts,
    color=['leiden', 'IRF9', 'RFX5', 'E2F4'],
    cmap='RdBu_r',
    size=1.5,
    vcenter=0,
    frameon=False
)
../_images/notebooks_spatial_39_0.png

Pathway activity inference

We can also infer pathway activities from our transcriptomics data.

PROGENy model

PROGENy is a comprehensive resource containing a curated collection of pathways and their target genes, with weights for each interaction. For this example we will use the human weights (other organisms are available) and we will use the top 500 responsive genes ranked by p-value. Here is a brief description of each pathway:

  • Androgen: involved in the growth and development of the male reproductive organs.

  • EGFR: regulates growth, survival, migration, apoptosis, proliferation, and differentiation in mammalian cells

  • Estrogen: promotes the growth and development of the female reproductive organs.

  • Hypoxia: promotes angiogenesis and metabolic reprogramming when O2 levels are low.

  • JAK-STAT: involved in immunity, cell division, cell death, and tumor formation.

  • MAPK: integrates external signals and promotes cell growth and proliferation.

  • NFkB: regulates immune response, cytokine production and cell survival.

  • p53: regulates cell cycle, apoptosis, DNA repair and tumor suppression.

  • PI3K: promotes growth and proliferation.

  • TGFb: involved in development, homeostasis, and repair of most tissues.

  • TNFa: mediates haematopoiesis, immune surveillance, tumour regression and protection from infection.

  • Trail: induces apoptosis.

  • VEGF: mediates angiogenesis, vascular permeability, and cell migration.

  • WNT: regulates organ morphogenesis during development and tissue repair.

To access it we can use decoupler.

[21]:
progeny = dc.get_progeny(organism='human', top=500)
progeny
Downloading annotations for all proteins from the following resources: `['PROGENy']`
[21]:
source target weight p_value
0 Androgen TMPRSS2 11.490631 0.000000e+00
1 Androgen NKX3-1 10.622551 2.242078e-44
2 Androgen MBOAT2 10.472733 4.624285e-44
3 Androgen KLK2 10.176186 1.944414e-40
4 Androgen SARG 11.386852 2.790209e-40
... ... ... ... ...
6995 p53 ZMYM4 -2.325752 1.522388e-06
6996 p53 CFDP1 -1.628168 1.526045e-06
6997 p53 VPS37D 2.309503 1.537098e-06
6998 p53 TEDC1 -2.274823 1.547037e-06
6999 p53 CCDC138 -3.205113 1.568160e-06

7000 rows × 4 columns

Activity inference with Multivariate Linear Model (MLM)

To infer pathway enrichment scores we will run the Multivariate Linear Model (mlm) method. For each spot in our slide (adata), it fits a linear model that predicts the observed gene expression based on all pathways’ Pathway-Gene interactions weights. Once fitted, the obtained t-values of the slopes are the scores. If it is positive, we interpret that the pathway is active and if it is negative we interpret that it is inactive.

4ec3c287a1f74e2ab5b842fd77bf9a87

We can run mlm with a simple one-liner:

[22]:
dc.run_mlm(
    mat=adata,
    net=progeny,
    source='source',
    target='target',
    weight='weight',
    verbose=True,
    use_raw=False
)

# Store in new obsm keys
adata.obsm['progeny_mlm_estimate'] = adata.obsm['mlm_estimate'].copy()
adata.obsm['progeny_mlm_pvals'] = adata.obsm['mlm_pvals'].copy()
Running mlm on mat with 4032 samples and 19813 targets for 14 sources.

The obtained scores (t-values)(mlm_estimate) and p-values (mlm_pvals) are stored in the .obsm key:

[23]:
adata.obsm['progeny_mlm_estimate']
[23]:
Androgen EGFR Estrogen Hypoxia JAK-STAT MAPK NFkB PI3K TGFb TNFa Trail VEGF WNT p53
AAACAAGTATCTCCCA-1 0.726582 2.053643 -1.000791 2.382140 6.987282 -2.039776 -0.701668 -3.040569 0.064893 3.832963 -2.758397 0.119536 0.258800 -4.506958
AAACAATCTACTAGCA-1 0.290291 1.083183 -1.330531 2.192247 7.869350 -2.282047 -0.815247 -3.588147 -0.337527 3.434620 -3.502797 0.230367 0.286089 -4.556519
AAACACCAATAACTGC-1 1.150137 2.426073 -1.008912 2.499938 7.776049 -2.682242 -0.782385 -3.192250 0.019723 3.800171 -3.115057 0.243773 0.280674 -4.311701
AAACAGAGCGACTCCT-1 0.953800 2.201080 -1.045912 1.849995 27.788458 -2.363654 -0.377379 -2.606445 -0.078141 3.240748 -3.111001 0.055545 0.199865 -5.400497
AAACAGCTTTCAGAAG-1 0.525047 1.662629 -1.394362 2.457380 6.697484 -2.214217 -1.591620 -3.209660 -0.590807 3.827924 -2.293088 -0.224760 0.807910 -5.429735
... ... ... ... ... ... ... ... ... ... ... ... ... ... ...
TTGTTTCACATCCAGG-1 0.749198 2.242120 -0.933600 2.753072 7.195352 -2.574513 -0.833506 -3.092716 -0.160780 3.545511 -2.567993 0.151230 0.009297 -4.479048
TTGTTTCATTAGTCTA-1 0.789263 2.098119 -1.414412 3.325075 7.465666 -2.504934 -0.525724 -3.662633 0.119874 3.946503 -2.805915 0.339055 0.133503 -4.287980
TTGTTTCCATACAACT-1 0.833514 1.983598 -1.009022 2.609354 6.795700 -2.645369 -0.830475 -2.940599 0.285104 3.994637 -2.749641 0.056184 0.261842 -4.687103
TTGTTTGTATTACACG-1 0.792543 2.018600 -1.130889 3.109919 8.298865 -2.645617 -0.529076 -3.662524 0.480168 3.827680 -2.748141 -0.197740 0.338115 -3.847667
TTGTTTGTGTAAATTC-1 0.868652 1.681782 -1.156972 2.128084 7.650702 -2.521203 -0.784943 -3.557293 -0.489564 3.589812 -3.457454 0.164759 0.512957 -4.315510

4032 rows × 14 columns

Visualization

Like in the previous section, we will extract the activities from the adata object.

[24]:
acts = dc.get_acts(adata, obsm_key='progeny_mlm_estimate')
acts
[24]:
AnnData object with n_obs × n_vars = 4032 × 14
    obs: 'in_tissue', 'array_row', 'array_col', 'n_genes', 'leiden', 'conn'
    uns: 'spatial', 'log1p', 'hvg', 'pca', 'neighbors', 'leiden', 'leiden_colors', 'dendrogram_leiden'
    obsm: 'spatial', 'X_pca', 'ulm_estimate', 'ulm_pvals', 'collectri_ulm_estimate', 'collectri_ulm_pvals', 'mlm_estimate', 'mlm_pvals', 'progeny_mlm_estimate', 'progeny_mlm_pvals'

Once extracted we can visualize them:

[25]:
sc.pl.spatial(
    acts,
    color=['Trail', 'leiden'],
    cmap='RdBu_r',
    vcenter=0,
    size=1.5,
    frameon=False
)
sc.pl.violin(
    acts,
    keys='Trail',
    groupby='leiden',
    rotation=90
)
../_images/notebooks_spatial_49_0.png
../_images/notebooks_spatial_49_1.png

Here we show the activity of the pathway Trail, which is associated with apoptosis.

Exploration

We can visualize which pathways are more active in each cluster:

[26]:
sc.pl.matrixplot(acts, var_names=acts.var_names, groupby='leiden', dendrogram=True, standard_scale='var',
                 colorbar_title='Z-scaled scores', cmap='RdBu_r')
../_images/notebooks_spatial_52_0.png

We can visualize again on the tissue some cluster-specific pathways:

[27]:
sc.pl.spatial(
    acts,
    color=['leiden', 'JAK-STAT', 'PI3K', 'TGFb'],
    cmap='RdBu_r',
    size=1.5,
    vcenter=0,
    frameon=False
)
../_images/notebooks_spatial_54_0.png

Functional enrichment of biological terms

Finally, we can also infer activities for general biological terms or processes.

MSigDB gene sets

The Molecular Signatures Database (MSigDB) is a resource containing a collection of gene sets annotated to different biological processes.

[28]:
msigdb = dc.get_resource('MSigDB')
msigdb
Downloading annotations for all proteins from the following resources: `['MSigDB']`
[28]:
genesymbol collection geneset
0 MAFF chemical_and_genetic_perturbations BOYAULT_LIVER_CANCER_SUBCLASS_G56_DN
1 MAFF chemical_and_genetic_perturbations ELVIDGE_HYPOXIA_UP
2 MAFF chemical_and_genetic_perturbations NUYTTEN_NIPP1_TARGETS_DN
3 MAFF immunesigdb GSE17721_POLYIC_VS_GARDIQUIMOD_4H_BMDC_DN
4 MAFF chemical_and_genetic_perturbations SCHAEFFER_PROSTATE_DEVELOPMENT_12HR_UP
... ... ... ...
3838543 PRAMEF22 go_biological_process GOBP_POSITIVE_REGULATION_OF_CELL_POPULATION_PR...
3838544 PRAMEF22 go_biological_process GOBP_APOPTOTIC_PROCESS
3838545 PRAMEF22 go_biological_process GOBP_REGULATION_OF_CELL_DEATH
3838546 PRAMEF22 go_biological_process GOBP_NEGATIVE_REGULATION_OF_DEVELOPMENTAL_PROCESS
3838547 PRAMEF22 go_biological_process GOBP_NEGATIVE_REGULATION_OF_CELL_DEATH

3838548 rows × 3 columns

As an example, we will use the hallmark gene sets, but we could have used any other.

Note

To see what other collections are available in MSigDB, type: msigdb['collection'].unique().

We can filter by for hallmark:

[29]:
# Filter by hallmark
msigdb = msigdb[msigdb['collection']=='hallmark']

# Remove duplicated entries
msigdb = msigdb[~msigdb.duplicated(['geneset', 'genesymbol'])]

# Rename
msigdb.loc[:, 'geneset'] = [name.split('HALLMARK_')[1] for name in msigdb['geneset']]

msigdb
[29]:
genesymbol collection geneset
233 MAFF hallmark IL2_STAT5_SIGNALING
250 MAFF hallmark COAGULATION
270 MAFF hallmark HYPOXIA
373 MAFF hallmark TNFA_SIGNALING_VIA_NFKB
377 MAFF hallmark COMPLEMENT
... ... ... ...
1449668 STXBP1 hallmark PANCREAS_BETA_CELLS
1450315 ELP4 hallmark PANCREAS_BETA_CELLS
1450526 GCG hallmark PANCREAS_BETA_CELLS
1450731 PCSK2 hallmark PANCREAS_BETA_CELLS
1450916 PAX6 hallmark PANCREAS_BETA_CELLS

7318 rows × 3 columns

Enrichment with Over Representation Analysis (ORA)

To infer functional enrichment scores we will run the Over Representation Analysis (ora) method. As input data it accepts an expression matrix (decoupler.run_ora) or the results of differential expression analysis (decoupler.run_ora_df). For the former, by default the top 5% of expressed genes by sample are selected as the set of interest (S), and for the latter a user-defined significance filtering can be used. Once we have S, it builds a contingency table using set operations for each set stored in the gene set resource being used (net). Using the contingency table, ora performs a one-sided Fisher exact test to test for significance of overlap between sets. The final score is obtained by log-transforming the obtained p-values, meaning that higher values are more significant.

f843e593825a4707be033be3e66e8d7e

We can run ora with a simple one-liner:

[30]:
dc.run_ora(
    mat=adata,
    net=msigdb,
    source='geneset',
    target='genesymbol',
    verbose=True,
    use_raw=False
)

# Store in a different key
adata.obsm['msigdb_ora_estimate'] = adata.obsm['ora_estimate'].copy()
adata.obsm['msigdb_ora_pvals'] = adata.obsm['ora_pvals'].copy()
Running ora on mat with 4032 samples and 19813 targets for 50 sources.
100%|████████████████████████████████████████████████████████████████████████████████████████████████████████████████████| 4032/4032 [00:08<00:00, 474.32it/s]

The obtained scores (-log10(p-value))(ora_estimate) and p-values (ora_pvals) are stored in the .obsm key:

[31]:
adata.obsm['msigdb_ora_estimate'].iloc[:, 0:5]
[31]:
source ADIPOGENESIS ALLOGRAFT_REJECTION ANDROGEN_RESPONSE ANGIOGENESIS APICAL_JUNCTION
AAACAAGTATCTCCCA-1 2.853878 27.778801 3.050313 0.626249 4.373918
AAACAATCTACTAGCA-1 1.824514 28.717391 1.688825 0.626249 3.082310
AAACACCAATAACTGC-1 3.242752 25.935842 2.558473 0.626249 3.491148
AAACAGAGCGACTCCT-1 1.824514 31.600349 3.050313 0.294479 1.994681
AAACAGCTTTCAGAAG-1 3.242752 21.533596 2.558473 0.294479 3.082310
... ... ... ... ... ...
TTGTTTCACATCCAGG-1 2.853878 21.533596 3.050313 0.626249 1.994681
TTGTTTCATTAGTCTA-1 3.653318 26.851581 4.138750 0.294479 3.082310
TTGTTTCCATACAACT-1 4.084916 29.667245 2.558473 0.626249 2.696111
TTGTTTGTATTACACG-1 2.144016 29.667245 2.103961 0.626249 3.082310
TTGTTTGTGTAAATTC-1 1.824514 26.851581 3.577630 0.626249 3.082310

4032 rows × 5 columns

Visualization

Like in the previous sections, we will extract the activities from the adata object.

[32]:
acts = dc.get_acts(adata, obsm_key='msigdb_ora_estimate')

# We need to remove inf and set them to the maximum value observed
acts_v = acts.X.ravel()
max_e = np.nanmax(acts_v[np.isfinite(acts_v)])
acts.X[~np.isfinite(acts.X)] = max_e

acts
[32]:
AnnData object with n_obs × n_vars = 4032 × 50
    obs: 'in_tissue', 'array_row', 'array_col', 'n_genes', 'leiden', 'conn'
    uns: 'spatial', 'log1p', 'hvg', 'pca', 'neighbors', 'leiden', 'leiden_colors', 'dendrogram_leiden'
    obsm: 'spatial', 'X_pca', 'ulm_estimate', 'ulm_pvals', 'collectri_ulm_estimate', 'collectri_ulm_pvals', 'mlm_estimate', 'mlm_pvals', 'progeny_mlm_estimate', 'progeny_mlm_pvals', 'ora_estimate', 'ora_pvals', 'msigdb_ora_estimate', 'msigdb_ora_pvals'

Once extracted we can visualize them:

[33]:
sc.pl.spatial(
    acts,
    color=['APICAL_JUNCTION', 'leiden'],
    cmap='RdBu_r',
    size=1.5,
    frameon=False
)
sc.pl.violin(
    acts,
    keys='APICAL_JUNCTION',
    groupby='leiden',
    rotation=90
)
../_images/notebooks_spatial_66_0.png
../_images/notebooks_spatial_66_1.png

Exploration

Let’s identify which are the top terms per cluster. We can do it by using the function dc.rank_sources_groups, as shown before.

[34]:
df = dc.rank_sources_groups(acts, groupby='leiden', reference='rest', method='t-test_overestim_var')
df
[34]:
group reference names statistic meanchange pvals pvals_adj
0 Clust. 0 rest UNFOLDED_PROTEIN_RESPONSE 11.159865 0.698623 9.250865e-28 4.625432e-26
1 Clust. 0 rest UV_RESPONSE_UP 9.698502 0.475151 1.473016e-21 1.841270e-20
2 Clust. 0 rest HYPOXIA 8.087560 0.635435 1.403690e-15 1.169741e-14
3 Clust. 0 rest MTORC1_SIGNALING 7.894915 0.677990 5.825152e-15 4.160823e-14
4 Clust. 0 rest ANDROGEN_RESPONSE 7.856351 0.295970 7.809971e-15 4.881232e-14
... ... ... ... ... ... ... ...
495 Clust. 9 rest EPITHELIAL_MESENCHYMAL_TRANSITION -7.434615 -1.922915 6.866049e-12 3.120931e-11
496 Clust. 9 rest APICAL_JUNCTION -8.167614 -0.809646 8.057563e-14 4.284914e-13
497 Clust. 9 rest ESTROGEN_RESPONSE_LATE -8.230622 -0.314651 5.937254e-14 3.710784e-13
498 Clust. 9 rest MITOTIC_SPINDLE -8.510308 -0.636282 5.161806e-15 3.687005e-14
499 Clust. 9 rest TGF_BETA_SIGNALING -10.227684 -1.119047 4.436417e-20 5.545522e-19

500 rows × 7 columns

We can then extract the top 3 terms per cluster:

[35]:
n_top = 3
term_markers = df.groupby('group').head(n_top).groupby('group')['names'].apply(lambda x: list(x)).to_dict()
term_markers
[35]:
{'Clust. 0': ['UNFOLDED_PROTEIN_RESPONSE', 'UV_RESPONSE_UP', 'HYPOXIA'],
 'Clust. 1': ['NOTCH_SIGNALING',
  'IL2_STAT5_SIGNALING',
  'PI3K_AKT_MTOR_SIGNALING'],
 'Clust. 2': ['WNT_BETA_CATENIN_SIGNALING',
  'ALLOGRAFT_REJECTION',
  'APOPTOSIS'],
 'Clust. 3': ['IL2_STAT5_SIGNALING', 'APICAL_SURFACE', 'TGF_BETA_SIGNALING'],
 'Clust. 4': ['EPITHELIAL_MESENCHYMAL_TRANSITION',
  'HEDGEHOG_SIGNALING',
  'APICAL_JUNCTION'],
 'Clust. 5': ['E2F_TARGETS', 'G2M_CHECKPOINT', 'MYC_TARGETS_V1'],
 'Clust. 6': ['COAGULATION', 'TNFA_SIGNALING_VIA_NFKB', 'COMPLEMENT'],
 'Clust. 7': ['COAGULATION', 'KRAS_SIGNALING_UP', 'COMPLEMENT'],
 'Clust. 8': ['MYC_TARGETS_V1', 'E2F_TARGETS', 'G2M_CHECKPOINT'],
 'Clust. 9': ['INTERFERON_GAMMA_RESPONSE',
  'INTERFERON_ALPHA_RESPONSE',
  'KRAS_SIGNALING_DN']}

We can plot the obtained terms:

[36]:
sc.pl.matrixplot(acts, term_markers, 'leiden', dendrogram=True, standard_scale='var',
                 colorbar_title='Z-scaled scores', cmap='RdBu_r', swap_axes=True)
../_images/notebooks_spatial_72_0.png

We can visualize again on the tissue some cluster-specific terms:

[37]:
sc.pl.spatial(
    acts,
    color=['leiden', 'INTERFERON_ALPHA_RESPONSE', 'G2M_CHECKPOINT','TNFA_SIGNALING_VIA_NFKB'],
    cmap='RdBu_r',
    size=1.5,
    frameon=False
)
../_images/notebooks_spatial_74_0.png