Next Article in Journal
Perimenopause Decreases SERCA2a Activity in the Hearts of a Mouse Model of Ovarian Failure
Previous Article in Journal
Targets to Search for New Pharmacological Treatment in Idiopathic Parkinson’s Disease According to the Single-Neuron Degeneration Model
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Accurate Identification of Spatial Domain by Incorporating Global Spatial Proximity and Local Expression Proximity

1
State Key Laboratory of Ophthalmology, Zhongshan Ophthalmic Center, Sun Yat-sen University, Guangzhou 510060, China
2
Center for Precision Medicine, Sun Yat-sen University, Guangzhou 510080, China
*
Authors to whom correspondence should be addressed.
Biomolecules 2024, 14(6), 674; https://doi.org/10.3390/biom14060674
Submission received: 3 May 2024 / Revised: 1 June 2024 / Accepted: 7 June 2024 / Published: 9 June 2024
(This article belongs to the Section Bioinformatics and Systems Biology)

Abstract

:
Accurate identification of spatial domains is essential in the analysis of spatial transcriptomics data in order to elucidate tissue microenvironments and biological functions. However, existing methods only perform domain segmentation based on local or global spatial relationships between spots, resulting in an underutilization of spatial information. To this end, we propose SECE, a deep learning-based method that captures both local and global relationships among spots and aggregates their information using expression similarity and spatial similarity. We benchmarked SECE against eight state-of-the-art methods on six real spatial transcriptomics datasets spanning four different platforms. SECE consistently outperformed other methods in spatial domain identification accuracy. Moreover, SECE produced spatial embeddings that exhibited clearer patterns in low-dimensional visualizations and facilitated a more accurate trajectory inference.

1. Introduction

Spatial transcriptomics (ST) captures gene expression profiles with spatial information, providing novel insights into tissue molecular heterogeneity. Applying ST technology plays a crucial role in identifying cell–cell interactions and signaling pathways within the tissue microenvironment, and has enabled groundbreaking discoveries across fields such as neuroscience [1], developmental biology [2], and cancer biology [3]. A variety of ST platforms have been developed, with varying levels of throughput and resolution. Image-based ST platforms, including STARmap [4], seqFISH [5,6], seqFISH+ [7], MERFISH [8] and FISSEQ [9], provide highly accurate gene expression measurement at single-cell resolution but only for a limited number of targeted genes [10]. On the other hand, sequencing-based ST platforms, such as spatial transcriptomics [11] and its commercial version, 10× Genomics Visium; Slide-seqV2 [12]; HDST [13]; Seq-Scope [14] and Stereo-seq [15], can perform high-throughput sequencing on a genome-wide scale with increasing spatial resolution. The spatial resolution of sequencing-based technology continues to improve, with Seq-Scope and Stereo-seq capable of merging subcellular spots based on cellular location to achieve single-cell resolution.
In ST data, spatial domains refer to regions exhibiting consistent patterns in both gene expression and physical location, each with specific anatomical structures [1,16]. Accurately identifying spatial domains is crucial for various downstream analyses, including trajectory inference, cell type deconvolution and cell–cell communications, as well as their biological interpretation. Spatial domains are distinct from cell types, which have been extensively studied in single-cell data. Cell types can be obtained by clustering transcriptional information, and their spatial distribution patterns are uncertain. They may be spatially concentrated, as in the case of excitatory neurons, or discretely distributed, as in the case of astrocytes [17]. Spatial domains are continuous in space, so relying solely on gene expression is insufficient to capture them. It is critical to incorporate spatial information while accurately capturing expression information; this presents a new challenge.
Several methods have been developed to address this challenge. Many existing techniques utilize spatial location information to find neighboring spots for each spot and enhance the similarity between neighbors to ensure the spatial continuity of the domain. Among them, BayesSpace [18] and BASS [19] perform latent variable modeling of regional labels and use Bayesian methods for inference. SpaGCN [20] and STAGATE [21] employ graph convolutional networks and graph attention networks to aggregate neighbor information, respectively. GraphST [22], SpaceFlow [23] and conST [24] utilize self-supervised graph-embedding learning strategies. However, spatial adjacency relationships only represent local information, neglecting to consider the fact that global structure and patterns may result in a lack of comprehensive understanding of the data. In contrast, SpatialPCA [25] introduces a spatially aware dimension reduction method, one which leverages global spatial relationship by measuring the similarity of pairwise spots. Nonetheless, focusing solely on global similarities may overlook subtle spatial details. Moreover, the simultaneous capture of both local and global information, harnessing the advantages of each, remains an area needing further exploration. In addition, some existing methods have limited effectiveness in extracting gene expression features. They use scaled values of highly variable genes (HVG) (e.g., GraphST, SpatialPCA, STAGATE and SpaceFlow) or employ dimensionality reduction techniques like principal component analysis (PCA) (e.g., BASS, SpaGCN, BayesSpace and conST) for expression features. However, these features encounter difficulty in handling expression with high noise, a condition which is very common in barcode-based sequencing ST methods, particularly in high-resolution techniques such as Stereo-seq.
To this end, we developed SECE, an accurate method for identifying spatial domains. SECE first employs an autoencoder (AE) with statistical modeling to obtain gene expression features, which we call cell type-related embedding (CE). Then, it incorporates global and local spatial proximity with CE to learn spatial embedding (SE). Global proximity is quantified by physical distance, and it is thought that the pairwise similarity between spots decreases with longer spatial distances. Local proximity, on the other hand, is determined by expression similarity, aggregating neighbor information based on the similarity of gene expression between spots. SECE utilizes graph attention network (GAT) for SE learning; it aggregates local expression similarity through an attention mechanism while simultaneously constraining the global spatial similarity, using a Gaussian kernel function. Subsequently, by performing clustering on the SE, we can derive the spatial domain to which each spot belongs. SECE also facilitate downstream analyses like visualization and trajectory inference. We demonstrated SECE’s versatility across diverse ST platforms, including high-resolution methods like STARmap/Slide-seqV2/Stereo-seq and lower-resolution platforms like Visium. SECE’s accurate spatial representations in brain and tumor datasets highlight its ability to gain biological insights from complex ST data.

2. Materials and Methods

2.1. Architecture Overview

SECE is a versatile instrument for modeling ST data across resolutions, including subcellular, single-cell, near-single-cell and multicellular (Figure 1A). It takes the gene expression matrix and spatial coordinate matrix of ST data as input, and outputs spatial domains and embeddings for each spot. First, SECE uses an AE module with a count distribution assumption to compress the expression matrix into expression features. Next, it converts spatial coordinates into local and global position relationships, storing them in the adjacency matrix (ADM) and spatial similarity matrix (SSM), respectively. Then, the GAT module is utilized to balance the expression similarity of local neighbors and the global spatial similarity to obtain the SE (Figure 1B). Finally, spatial domains are identified by clustering SE using mclust [26]. Additionally, downstream analyses, including low-dimensional visualizations [27] and trajectory inference [28,29], are derived from the SE (Figure 1C).

2.2. Extracting Expression Features with AE Module

Given ST data with N spots and G genes, the dimensions of gene expression matrix X and spatial coordinate matrix Y are N × G and N × 2, respectively. The raw counts X are normalized by library size and then log-transformed to obtain the normalized expression matrix X ~ . We first employ AE with zero-inflated negative binomial (ZINB) or negative binomial (NB) distribution [30,31] to compress X ~ into low-dimensional features Z. Let xng denote the count value of gene g in spot n; the likelihood function of xng under ZINB and NB distributions is given by
Z I N B x n g ; π n g ,   r n g ,   p g = π n g δ 0 x n g + 1 π n g Γ x n g + r n g x n g ! Γ r n g p g r n g 1 p g x n g
and
N B x n g ; r n g , p g = Γ x n g + r n g x n g ! Γ r n g p g r n g ( 1 p g ) x n g ,
where δ 0 is the Dirac delta function, π n g is the probability of true gene expression being 0, Γ is the gamma function and (rng, pg) is the standard parametrization for the NB distribution.
The AE module takes X ~ as input, and outputs distribution parameters. The formulation is
Z = f e X ~
Z = f d 1 Z
N Π , R , P = f d 2 Z ,
where fe represents an encoder, while fd1 and fd2 constitute the decoders. Specifically, fe consists of two nonlinear layers, each utilizing Rectified Linear Unit (ReLU) activation functions. These layers reduce the feature dimension G into m′ and m, respectively, yielding the expression feature Z. Subsequently, fd1 decodes Z into Z′ with a feature dimension of m′. The expression fd2 comprises three output layers which take Z′ as input and output three parameter matrices Π , R , P of the ZINB distribution, each consisting of elements π n g , r n g , p g . The activation functions of these three output layers are exponential, sigmoid and exponential functions, respectively. The expression fd2 employs two layers to learn (R, P) under the NB distribution assumption.
The goal is to minimize the reconstructed loss by minimizing the negative log-likelihood (NLL) function, that is, L o s s p r e = N L L Z I N B X ; Π , R , P or Losspre = NllNB(X; R, P). We employed ZINB for highly sparse data like Stereo-seq and Slide-seq, and NB for less sparse data, including STARmap and Visium. Both are implemented in the SECE package.

2.3. Capturing Local and Global Relationships

To incorporate physical location information, we construct an ADM and SSM from the spatial coordinates Y. ADM summarizes local spatial relationships by storing neighbors for each spot. This local neighborhood information is later employed to adaptively aggregate features in a GAT module based on expression similarity. In contrast, SSM captures global spatial proximity by providing a spatial similarity measure for all pairs of spots, not just neighbors. The SSM is subsequently utilized to constrain the global similarity of SE.
ADM A is a N × N-dimensional symmetric matrix where elements are assigned values of 1 or 0 to indicate neighboring spots or non-neighboring spots, respectively. More precisely, the element Aij denotes the adjacent relationship between spot i and spot j:
A i j = 0 ,                               v i N j                 1 ,                               v i N j .            
Here, N j represents the set of spatial neighbors of spot j, which can be determined based on coordinates Y by employing K-Nearest Neighbor (KNN) or applying a distance cutoff. By default, we utilize KNN, with the number of neighbors set to 6 for Visium datasets and 10 for other datasets.
SSM Σ is also a N × N-dimensional symmetric matrix, wherein elements decrease as the distance between spots increases, exhibiting an exponential decay tendency. For spot i and spot j with coordinates y i = ( y i 1 ,   y i 2 ) and y j = ( y j 1 ,   y j 2 ) , the corresponding element is:
Σ i j = e x p y i y j 2 2 γ
where the bandwidth parameter γ controls the spatial influence. By default, γ is set as the 0.05 quantile distance. A larger γ results in a greater spatial influence.

2.4. Learning SE with GAT Module

After capturing the expression features and extracting local and global position relationships, we employ the GAT module to learn SE, subsequently clustering SE to delineate spatial domains. The GAT module consists of two GAT layers.
We first introduce the GAT layer. It takes feature matrix and ADM as inputs, and outputs a new feature matrix after aggregating neighbor information. Let H = (h1, h2, …, hN) denote the input feature matrix, which has dimensions N × m0, with N samples and m0 features. The output of GAT layer is denoted as H = ( h 1 ,   h 2 ,   ,   h N ) with dimensions N × m 0 . The GAT layer performs aggregation for each sample adaptively based on the normalized attention scores. For sample j, the output feature h j can be formulated as follows:
h j = σ i N j α i j W h i
where W is a weight matrix with dimensions m 0 × m 0 , N j represents the set of neighboring samples of sample j and αij is the normalized attention coefficient matrix using the SoftMax function:
α i j = s o f t m a x i e i j = exp e i j k N j exp e k j
where e i j = a T W h i W h j , a is learnable vector and is the concatenation operation. We used the Exponential Linear Unit (ELU) as activation function σ in the GAT layer.
The GAT module in SECE consists of two GAT layers. It takes expression features Z and ADM A as input, and outputs SE matrix U, which is a N × m-dimensional matrix:
U = G A T 2 G A T 1 Z , A , A
During neighbor information aggregation using GAT, local expression similarities are captured via attention and adaptively aggregated. To preserve as much information as possible in expression features, the local learning target is the reconstruction loss Llocal = MSE(U, Z). We further constrain pairwise correlation using SSM, which including global information, that is, the correlation of each SE at N positions UUT is close to Σ , L g l o b a l = M S E U U T , Σ . The objective function balances the two similarities by λ g l o b a l and λ l o c a l :
L o s s = λ g l o b a l L g l o b a l + λ l o c a l L l o c a l
We kept λ l o c a l = 1 unchanged, and adjusted λ g l o b a l based on the ST platform. After obtaining the final SE U, we utilized the clustering method mclust [26] to cluster the U and determine the spatial domain for each spot.

2.5. Architecture of SECE

In the AE module, the dimension of the hidden layer m′ and bottleneck layer m are 128 and 32, respectively. An Adaptive Moment Estimation (Adam) optimizer is used to minimize Losspre, with a learning rate of 1 × 10−3 and dropout rate of 0.1. For the GAT module, the dimensions of the two GAT layers are both 32. The Adam optimizer is employed to minimize Loss, with a learning rate of 1 × 10−2 and dropout rate of 0.2. The default number of iterations for the AE and GAT modules are set to 40 and 50, respectively.
The hyperparameters λ g l o b a l and λ l o c a l control the contributions of global and local similarities, respectively, and a larger λ g l o b a l gives greater global influence. Each ST platform has different resolutions and fields of view. For example, spots of ST arrays contain dozens of cells, while Stereo-seq only contains a single cell. Stereo-seq can sequence the hemibrain, while STARmap can only detect a minor region of the visual cortex. We choose a smaller λ g l o b a l value for platforms with more spots and higher resolution. Specifically, for Stereo-seq and Slide-seqV2 data including over 10,000 spots, and with approximately single-cell resolution, we used λ g l o b a l = 0.08 . For Visium data with several thousand spots, λ g l o b a l was set to 0.3. For STARmap data with only 1207 cells, we set λ g l o b a l to 2. For datasets from other platforms, users could also select λ g l o b a l while following this standard.

2.6. Datasets

A mouse visual cortex [4] was generated from a STARmap platform with 1207 cells, 1020 genes and a sparsity of 76.88%. STARmap is an in situ sequencing-based ST method with single-molecule resolution. Despite its low gene throughput (160 to 1020 genes), it offers high sensitivity at single-cell resolution, with high efficiency and reproducibility. A mouse hippocampus dataset was generated from Slide-seqV2 [12] platform, with 53,208 cells and 23,264 genes from hippocampus, cortex and thalamus, boasting a high sparsity of 98.19%. Slide-seqV2 offers transcriptome-wide sequencing with near-cellular resolution (10 μm). A mouse olfactory bulb [32] was generated from the Stereo-seq [15] platform, comprising 19,527 cells and 27,106 genes, while 98.69% of values were zero. Stereo-seq is an emerging technique for ST with genome-wide throughput and subcellular resolution. This method captures the expression profile and spatial coordination of each DNA nanoball (DNB) and employs image-based cell segmentation to segment single cells. A mouse hemibrain was also generated from Stereo-seq [15], one which contained 50,140 cells and 25,879 genes, with 96.94% of the values being 0. Human breast cancer data was generated from the Visium platform, containing 3798 spots and 24,923 genes, of which only 77.44% were zero values. Visium is the commercial version of Spatial transcriptomics [11], with a low resolution of 55 μm spots and 1–10 cells per spot [33]. A human dorsolateral prefrontal cortex (DLPFC) [1] dataset was also generated from the Visium platform. In total, 12 DLPFC slices were annotated manually and used as the ground truth of the spatial domain identification.
For Stereo-seq datasets, cells with an expression level below 200 were removed, according to procedures used in the original studies [15]; then, we filtered the genes expressed in fewer than 20 cells. For other datasets, we screened spots with an expression level below 20 and genes that expressed less than 20 spots. The filtered expression matrix and its corresponding coordinates were input into SECE for analysis.
For datasets with manual annotation, such as STARmap cortex, DLPFC and breast cancer data, we select the number of domains according to their original study. For mouse hippocampus, hemibrain and olfactory bulb data, we determined the numbers based on the ABA organizational structure.

2.7. Evaluation Metrics

The ARI evaluates the degree of overlap between the two divisions, which is formulated as
A R I = i j n i j 2 [ i a i 2 j b j 2 ] / N 2 1 2 [ i a i 2 + j b j 2 ] [ i a i 2 j b j 2 ] / N 2 ,
where N is the number of samples and nij, ai and bj are values from the contingency table. Specifically, ai represents the number of samples with the real category label i; bj represents the number of samples with the predicted label j; and nij represents the number of samples with the real category label i and the predicted label j. In this paper, we utilize ARI to evaluate the consistency of the spatial domains identified by various methods with the ground truth domains. The ARI ranges from −1 to 1; a greater value indicates better agreement with the true labels.
The ACC evaluates the correctness of categories, which is calculated as
A C C = i = 1 N δ ( r i ,   o ( s i ) ) N ,
where N is the number of samples, and ri and si are the true and predicted spatial domain label of spot i. δ is a function that can be defined as
δ x , y = 1 ,                               x = y 0 ,                               x y
o is a mapping function that takes the real label ri as the reference label and then rearranges si in the same arrangement, which is implemented using the classical Kuhn–Munkres algorithm [34]. The ACC ranges from 0 to 1, and a greater value indicates better performance.
The ASW describes the degree of match between features and category labels. For every spot i, silhouette width S(i) is calculated as
S i = b i a ( i ) m a x { a i , b ( i ) }
where a(i) is the average distance between i and the points in its cluster, and b(i) is the lowest average distance from i to points in other clusters. In this paper, we use ASW to evaluate how well the SE obtained by various methods explain the known spatial layers. Distance is calculated by SE of various methods, and clusters are annotated spatial layers. The ASW ranges from −1 to 1. A greater ASW indicates better SE learning.
The LISI [35] measures the degree of local mixing to evaluate the level of spatial aggregation patterns. For each spot i, LISI can be formulated as
L I S I i = 1 l L p i l
where p(l) is the probability that the spatial domain cluster label l exists in the local neighborhood of spot i, and L is the set of spatial domains. In this paper, we use LISI to evaluate the spatial aggregation degree of spatial domains. Local neighborhoods are generated by spatial location, and cluster labels are those predicted by each algorithm. LISI values ranges from 0 to 1. A smaller LISI indicates better spatial aggregation patterns, i.e., less mixing of cluster labels within local spatial neighborhoods.

2.8. Methods for Comparison

We compared SECE with the existing spatial domain identification methods: (1) BayesSpace, implemented in the R package BayesSpace V1.4.1 downloaded from https://github.com/edward130603/BayesSpace (accessed on 23 April 2022); (2) SpaGCN, implemented in the Python package SpaGCN V1.2.2 downloaded from https://github.com/jianhuupenn/SpaGCN (accessed on 23 April 2022); (3) STAGATE, implemented in the Python package STAGATE_pyG V1.0.0 downloaded from https://github.com/QIFEIDKN/STAGATE_pyG (accessed on 23 April 2022); (4) BASS, implemented in the R package BASS V1.3.1 from https://github.com/zhengli09/BASS (accessed on 3 June 2023); (5) SpaceFlow, implemented in the Python package SpaceFlow V1.0.4 from https://github.com/hongleir/SpaceFlow (accessed on 3 June 2023); (6) GraphST, implemented in the Python package GraphST V1.4.1 from https://github.com/JinmiaoChenLab/GraphST (accessed on 3 June 2023); (7) SpatialPCA, implemented in the R package SpatialPCA V1.2.0 from https://github.com/shangll123/SpatialPCA (accessed on 3 June 2023); and (8) conST (we referred to https://github.com/ys-zong/conST, accessed on 3 June 2023, to run conST).
We ran each method with its default parameters. For methods that can output spatial features, we used their default feature dimensions. Specifically, the dimensions of GraphST, SpatialPCA, STAGATE and SpaceFlow were 20, 20, 30 and 50 dimensions, respectively. These SEs were used to compute ASW, generate UMAP low-dimensional visualization, and perform trajectory inference. UMAP and domain-level trajectory inference PAGA [28] were computed using ‘scanpy.tl.umap’ and ‘scanpy.tl.paga’ from the scanpy V1.9.3 package, respectively, while cell-level trajectory inference followed Monocle3 V1.0.0 [29] guidelines https://cole-trapnell-lab.github.io/monocle3/ (accessed on 14 May 2022).

3. Results

3.1. Application to STARmap Data

We first tested SECE on the mouse visual cortex data generated by STARmap [4], with ground truth layers. The 1207 cells were divided into seven layers, including Layer (L)1, L2/3, L4, L5 and L6, as well as the corpus callosum (CC) and hippocampus (HPC) (Figure 2A).
First, we compared the spatial domain identification results of SECE with eight existing methods, including BASS, SpaceFlow, GraphST, STAGATE, SpatialPCA, SpaGCN, BayesSpace and conST (Figure 2B,C). The consistency values for spatial domains and ground truth layers were evaluated using the Adjusted Rand Index (ARI) and Accuracy (ACC) based on Kuhn–Munkres [34]. SECE achieved the highest consistency, with an ARI value of 0.65 and an ACC value of 0.79 (Figure 2D). It was closely followed by BASS, with ARI of 0.63, however, BASS incorrectly combined HPC and L5 (domain 5), which prevents the utilization of the Kuhn–Munkres algorithm to reassign clusters and compute the accuracy (ACC). Furthermore, other algorithms also exhibited weak spatial aggregation, as indicated by the local inverse Simpson’s index (LISI) [35] (Figure S1A). Additionally, these methods mistakenly classified aggregated endothelial cells (Endo) in L1 and L2/3 into the same domain, as seen in domain 2 in SpaceFlow, domain 3 in STAGATE and BayesSpace, domain 5 in GraphST and SpatialPCA and domain 6 in SpaGCN (Figure 2C and Figure S1B).
Next, we compared the SE of SECE with those of SpaceFlow, GraphST, STAGATE and SpatialPCA. Due to the difficulty encountered by conST in generating clear spatial domains, its SE learning comparison was excluded. The SE of SECE explained the ground truth layers most effectively (ASW = 0.16), followed by SpaceFlow (ASW = 0.12), GraphST (ASW = 0.07), STAGATE (ASW = 0.07) and SpatialPCA (ASW = 0.05) (Figure 2D, right). Moreover, SECE had a clearer and continuous pattern in UMAP visualization, compared to the other methods (Figure 2E). When comparing trajectory inference results, we selected the cortex part in ground truth, namely, L1, L2/3, L4, L5 and L6. The PAGA [28] analysis based on SECE revealed a linear and continuous relationship between these layers (Figure 2F). Conversely, SpaceFlow mistakenly made the connection between L1 versus L4 and L5, while the patterns in the other three methods were more unclear. For individual cells, SECE exhibited a sequential increase of pseudo-time from L6 to L1 (Figure S1C,D) based on Monocle3 [29], and a similar trend was also observed in SpaceFlow. However, GraphST, STAGATE and SpatialPCA ordered L1 and L2/3 incorrectly (Figure S1E,F).

3.2. Application to Slide-seqV2 Data

We further tested SECE on the hippocampal dataset generated by the Slide-seqV2 platform [12]. SECE identified 14 distinct domains, and we annotated them according to the known structure of Allen Brain Atlas (ABA) (Figure 3A). The domains were hippocampus (Cornu Ammonis (CA)1, CA2, CA3, Dentate gyrus (DG) and CA slm/so/sr); cortex (Layers 4, 5a, 5b and 6); third ventricle; CC; and three subregions of the thalamus (Figure 3B, left). We further meticulously verified the subtle spatial domains. For example, the four important components of the hippocampal region, CA1, CA2, CA3 and DG, were clearly demarcated, as evidenced by the high expression of their known markers Wsf1, Rgs14, Nptxr and C1ql2 [36], respectively (Figure 3C and Figure S2A). Layer 5 in cortical regions was identified as two sublayers, Layer 5a and 5b, with different gene expression levels (Figure S2A,B). The composition of cell types in each domain further supported the delineation (Figure S3).
For comparison, we evaluated the performance of existing methods for spatial domain identification (Figure 3D). Notably, SECE was the only approach that could accurately detect subregions in both the cortex and CA. Specifically, for cortical areas, BASS failed to distinguish between Layer 4 and Layer 5, SpaceFlow mixed Layer 5 and Layer 6, SpatialPCA exhibited suboptimal division smoothness, and the remaining algorithms were unable to generate clear cortical regions. In the case of CA, except for GraphST, none of the methods succeeded in identifying the CA2 region. Furthermore, SECE exhibited the highest spatial aggregation performance, as it had the smallest LISI values (Figure S4A).
We also compared the performance of SE. UMAP generated from SECE clearly displayed clustering patterns for the hippocampus, cortex, thalamus and third ventricle, as well as their subregions (Figure 3B, right). In addition, their sublayers, like Layer 4 and Layer 6, were arranged in a sequence, while UMAP of STAGATE and GraphST missed them (Figure S4B). Moreover, we conducted trajectory inference for the cortex due to its continuous relationship between sublayers. We selected the domains corresponding to the cortex in each method and started with the deepest clusters, such as Layer 6 for SECE, domain 1 for SpaceFlow, domain 2 for GraphST and domain 3 for SpatialPCA and STAGATE (Figure 3E,F). For SECE, the pseudo-time consistently increased with decreasing cortical depth, clearly capturing the pseudo-time relationship between different layers. In contrast, SpaceFlow exhibited fewer distinguishable differences between the learned layers. SpatialPCA reversed the relationship between Layer 5b and Layer 6, while GraphST and STAGATE failed to identify the relationships within the cortical region.

3.3. Application to Stereo-Seq Data

In this section, we assessed on the mouse olfactory bulb [32] Stereo-seq [15] data. Spatial domains identified by SECE were annotated based on known olfactory bulb layers, including rostral migratory stream (RMS), granule cell layer (GCL), inner plexiform layer (IPL), mitral cell layer (MCL), external plexiform layer (EPL), glomerular layer (GL) and olfactory nerve layer (ONL), from inside to outside (Figure 4A and Figure S5A). These domains were validated using known markers [36] for each layer (Figure 4B). Notably, besides the known seven layers, we made a finer division of GCL and ONL, and the sublayers were named GCL-Inner, GCL-Outer, ONL-Inner and ONL-Outer, respectively. Several points of evidence confirmed these sublayers. First, marker gene expression differed between them; specifically, the GCL-inner highly expressed Nrgn and the GCL-outer highly expressed Pcp4. Markers in the ONL-outer also exhibited higher expression levels compared to those in the ONL-inner. Second, sublayers exhibited distinct cell type composition (Figure S5B–D). The GCL-Outer was almost composed of GC, while GCL-Inner contained a certain amount of Oligo, and the GC subtypes in GCL-Inner and GCL-Outer also differed. Moreover, the ONL-Outer almost exclusively contained OEC, while the ONL-Inner contained a fraction of OEC and more Astro. This supported previous findings that the ONL, as a part of the olfactory bulb blood–brain barrier, had fine internal and external subregions, with different cell types [37]. Our study provided further support for this finding at the spatially resolved single-cell level.
We also compared the performance of SECE to the existing methods (Figure 4C). BASS failed to separate the RMS from the GCL layer and could not distinguish GL from EPL. It also identified some irrelevant domains with few cells, such as domains 5, 8 and 9. STAGATE and SpaGCN mixed GL and EPL. SpaceFlow, GraphST and SpatialPCA encountered challenges in accurately dividing the GCL layer. BayesSpace and conST did not yield clear domain identifications. Additionally, BASS had the highest spatial aggregation performance, as well as the lowest LISI value, while SECE ranked second (Figure S5E). Furthermore, SECE showed robustness when tested with different number of clusters (7, 8 and 10), consistently providing well-bounded ring stratification (Figure S6). Moreover, we assessed the SE learning capabilities. The UMAP visualization based on SECE exhibited a continuous low-dimensional pattern, with nine layers arranged in order of spatial position from inside to outside (Figure 4D). The trajectory inference for these layers demonstrated an approximately linear relationship (Figure 4E). SpatialPCA also displayed linear patterns, except for domains 6 and 7, while the results of SpaceFlow, GraphST and STAGATE exhibited many false positive connections between domains.
We further applied SECE to a mouse hemibrain dataset with a more complex anatomic structure [15]. SECE achieved spatial domain annotations that were in highly consistent with ABA anatomy (Figure 5A,B(left)). We could clearly separate different regions, including cortical regions, hippocampal regions, midbrain regions, thalamic regions and fiber tracts (FT). The cortex included five layers, L1, L2/3, L4, L5, L6, LVC and CAA, which were supported by their cell type composition (Figures S7 and S8). In contrast, other methods failed to accurately identify these cortical layers (Figure 5C and Figure S9). Specifically, all of them encountered difficulties in identifying Cortex L5 with high smoothness. In addition, LISI showed that spatial domains identified by SECE had the strongest spatial pattern (Figure S9A). Furthermore, low-dimensional visualization of SE using t-SNE [38] showed the effects of aggregation in the same region and separation in different regions (Figure 5B, right).

3.4. Application to Visium Data

Finally, we tested the applicability of SECE on Visium dataset. The breast cancer dataset was divided into four phenotypic regions according to pathological images: ductal carcinoma in situ/lobular carcinoma in situ (DCIS/LCIS), healthy tissue (Healthy), invasive ductal carcinoma (IDC) and tumor-surrounding regions with low features of malignancy (Tumor edge), as well as 20 annotated subdivisions [32] (Figure 6A and Figure S10A).
We initially segmented 20 domains using each algorithm and performed phenotype annotations, which were then compared with image-based manual annotations (Figure 6B and Figure S10B,C). Notably, in the context of a dataset characterized by a low missing rate of 77.44%, all algorithms demonstrated commendable performance. Most of the algorithms had ARI values ranging from 0.55 to 0.62 (Figure S10D). We found that SECE divided individual tumor regions into wrapped layers more clearly. Specifically, cluster 20 and cluster 15 divided the IDC-5 into internal and external layers, while cluster 11 and cluster 10 split the DCIS/LCIS-1. (Figure 6B and Figure S10A). To investigate the biological significance of the refined stratification, we named clusters 20,15,11 and 10 as IDC-inner, IDC-outer, DCIS/LCIS-inner and DCIS/LCIS-outer, respectively. We also focused on the parts of clusters 4 and 5 that were located near the tumor edge, referred to as IDC-edge and DCIS/LCIS-edge. (Figure 6C). We analyzed the cell type composition of each spot in these domains by integrating the annotated breast cancer scRNA-seq data [39] and deconvoluting each spot using cell2location [40] (Figures S11 and S12). The proportion of tumor cells gradually decreased as the edge of the tumor was approached, while those of immune cells and stromal cells gradually increased (Figure 6D), confirming the correctness of the subregions.
We further explored the characteristics of these subregions (-edge, -outer and -inner) of DCIS/LCIS and IDC (Figure 6E). Differential expression analysis [41] showed that DCIS/LCIS-edge had a high level of humoral immunity, which could be confirmed by the enrichment of B cell receptor signaling pathway-related genes (IGLC2, IGLC3, IGHG1, IGHG3 and IGHA1). The DCIS/LCIS-outer subregion overexpressed KRT14 of the keratin family, which is a key regulator of metastasis, suggesting invasive potential [42,43]. In contrast, DCIS/LCIS-inner subregion showed non-metastatic traits but also had tumor-promoting abilities, as indicated by a high expression of LDHA [44]. As for the IDC subregions, IDC-edge showed elevated expression of biomarkers linked to tumor proliferation, invasion and migration, including tumor-associated macrophage (APOE), complement components (C1qA, C1qB), cathepsin (CTSD), and apolipoprotein (APOC1) [45,46,47,48,49]. IDC-outer had a high level of immunity and some transferability, derived from the high expression of MHC class I-related genes (HLA-A, HLA-B and B2M) [50,51]. Besides, there was increased expression of genes such as MGST1 and MRPS30-DT, which have been known to promote breast carcinoma cell growth and metastasis [52,53]. In IDC-inner, there were higher levels of tumor activity and lower levels of immune response. LINC00052, known to promote breast cancer cell proliferation by increasing signals of epidermal growth factor receptor (EGFR) such as HER3 [54,55,56], was overexpressed. Upregulation of COX6C and FAM234B implied higher levels of cellular respiration [57] and lower immune response function [58], respectively.
Furthermore, we inferred developmental trajectories for the IDC and DCIS/LCIS regions. The starting points for calculating pseudo-time in IDC and DCIS/LCIS were the interior of the tumor regions (Figure 6F). There were increased pseudo-time values as the region of the tumor moved outward, effectively mimicking the gradual progression of tumor development.
Moreover, to further evaluate the power of SECE in spatial domain identification, we tested 12 human dorsolateral prefrontal cortex (DLPFC) datasets generated from the Visium platform. The original study [1] had manually annotated the spatial domains of these datasets, encompassing white matter and six cortical layers (Figure 7A and Figure S13). The spatial domain identified by SECE exhibited the highest levels of agreement with the original annotations (Figure 7B,C and Figure S13). The median ARI for SECE was 0.58, surpassing the second- and third-ranked algorithms, STAGATE and SpatialPCA, which achieved median ARI values of 0.55 and 0.54, respectively (Figure 7D). These findings highlighted the superior performance of SECE in accurately delineating the spatial domains within the low-resolution data.

4. Discussion

SECE captures both local and global relationships among spots and aggregates their information using expression similarity and spatial similarity, respectively. This approach enables precise spatial domain division and facilitates interpretable spatial embedding learning across diverse ST datasets. Moreover, the AE module that explicitly models gene expression counts enhances SECE’s ability to handle noisy data.
With the increases in captured area within the ST data and advancements in resolution, there is a growing demand for computational methods which can be used to exhibit higher efficiency and scalability. We recorded the runtime and GPU memory consumption for each dataset (Figure S14). For Slide-seqV2 hippocampus data and the Stereo-seq hemibrain data, which contained over 50,000 cells, SECE achieved a running time of less than 4.2 min while utilizing less than 5GB of GPU memory. These results demonstrated the superior computational efficiency and scalability of SECE when dealing with large-scale datasets.
While SECE has demonstrated notable performance, there are still several aspects that can be further enhanced. Firstly, we employ a pre-defined SSM to characterize global similarity, but exploring more flexible global correlation patterns could be advantageous. Secondly, we only utilized ST data as inputs, but incorporating matching histology data may provide additional benefits [20]. Although matching histology image data is currently only available on specific platforms like Visium, we can still leverage such images as optional supplementary information when available. Finally, integrating single-cell data with ST data can enhance the data quality of the latter, increasing the throughput, or reducing the noise in the gene expression [59]. Therefore, incorporating single-cell data is another approach which can be used to improve spatial representation capabilities.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/biom14060674/s1, Figure S1: Trajectory inference on mouse visual cortex STARmap data, related to Figure 2; Figure S2: Marker of spatial domains identified by SECE on mouse hippocampus Slide-seqV2 data, related to Figure 3; Figure S3: Cell type composition of domains in mouse hippocampus Slide-seqV2 data, related to Figure 3; Figure S4: Spatial domains of mouse hippocampus Slide-seqV2 data, related to Figure 3; Figure S5: Spatial domain identification of olfactory bulb, related to Figure 4; Figure S6: Different numbers (7, 8, 9, 10) of spatial domains identified by SECE, STAGATE, SpaGCN and BayesSpace in the Stereo-seq mouse olfactory bulb data; Figure S7: Relationships between cell types and spatial regions identified by SECE of mouse brain Stereo-seq data, related to Figure 5; Figure S8: Gene expression heatmaps of cell type clusters in mouse brain Stereo-seq data, related to Figure 5; Figure S9: Spatial domain identification of mouse brain Stereo-seq data, related to Figure 5; Figure S10: Spatial domain identification of breast cancer data, related to Figure 6; Figure S11: Cell type composition in spatial regions identified by SECE, related to Figure 6; Figure S12: Number of cells per spot for each cell type inferred by cell2location in human breast cancer Visium data; Figure S13: Spatial domains identified by SECE, BASS, SpaceFlow, GraphST, STAGATE, SpatialPCA, SpaGCN, BayesSpace, and conST and manual annotation in 12 sections of the DLPFC dataset; Figure S14: Datasets used by SECE and their running information.

Author Contributions

Z.X. conceived and supervised the study. Y.Y. and Z.X. designed the study. Y.Y. analyzed the data. All of the authors interpreted the data. Y.Y. and Z.X. wrote the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Key Research and Development Program of China, grant number 2019YFA0904400, and the Key-Area Research and Development Program of Guangdong Province, grant number 2023B1111020006.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Slide-seq datasets are available at https://portals.broadinstitute.org/single_cell/study/slide-seq-study (accessed on 10 April 2022). The Stereo-seq mouse hemibrain dataset is available at https://db.cngb.org/stomics/mosta/ (accessed on 25 April 2022). The Stereo-seq olfactory bulb dataset is available at https://github.com/JinmiaoChenLab/SEDR_analyses (accessed on 28 March 2022). The Slide-seqV2 hippocampus datasets are available at https://singlecell.broadinstitute.org/ (accessed on 4 May 2022). The STARmap mouse visual cortex dataset is available at http://clarityresourcecenter.org/ (accessed on 4 May 2022). The Visium human breast cancer dataset is available at https://www.10xgenomics.com/resources/datasets/human-breast-cancer-block-a-section-1-1-standard-1-1-0 (accessed on 8 August 2022). The Visium DLPFC dataset is available within the spatialLIBD package (http://spatial.libd.org/spatialLIBD, accessed on 22 June 2023). We also organized these raw data, including expression counts and coordinates, as well as H and E images of Visium, into a format that is easy to read by SCANPY; the results are available at https://drive.google.com/drive/folders/1uHc2F_e1PX1Q_efuO5xrFw9bhJa0wCm4. The SECE algorithm is implemented and provided as a pip installable Python package, which is available on Github https://github.com/yuyuanyuana/SECE. The source code and datasets are available at Zenodo https://zenodo.org/record/8130682.

Acknowledgments

We acknowledge the support from Youjin Hu and thank all the data and software contributors who made this research possible. We also appreciate the longstanding support from the Center for Precision Medicine at Sun Yat-sen University.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Maynard, K.R.; Collado-Torres, L.; Weber, L.M.; Uytingco, C.; Barry, B.K.; Williams, S.R.; Catallini, J.L., 2nd; Tran, M.N.; Besich, Z.; Tippani, M.; et al. Transcriptome-scale spatial gene expression in the human dorsolateral prefrontal cortex. Nat. Neurosci. 2021, 24, 425–436. [Google Scholar] [CrossRef] [PubMed]
  2. Wang, M.; Hu, Q.; Lv, T.; Wang, Y.; Lan, Q.; Xiang, R.; Tu, Z.; Wei, Y.; Han, K.; Shi, C.; et al. High-resolution 3D spatiotemporal transcriptomic maps of developing Drosophila embryos and larvae. Dev. Cell 2022, 57, 1271–1283.e4. [Google Scholar] [CrossRef] [PubMed]
  3. Hunter, M.V.; Moncada, R.; Weiss, J.M.; Yanai, I.; White, R.M. Spatially resolved transcriptomics reveals the architecture of the tumor-microenvironment interface. Nat. Commun. 2021, 12, 6278. [Google Scholar] [CrossRef]
  4. Wang, X.; Allen, W.E.; Wright, M.A.; Sylwestrak, E.L.; Samusik, N.; Vesuna, S.; Evans, K.; Liu, C.; Ramakrishnan, C.; Liu, J.; et al. Three-dimensional intact-tissue sequencing of single-cell transcriptional states. Science 2018, 361, eaat5691. [Google Scholar] [CrossRef] [PubMed]
  5. Lubeck, E.; Coskun, A.F.; Zhiyentayev, T.; Ahmad, M.; Cai, L. Single-cell in situ RNA profiling by sequential hybridization. Nat. Methods 2014, 11, 360–361. [Google Scholar] [CrossRef] [PubMed]
  6. Shah, S.; Lubeck, E.; Zhou, W.; Cai, L. In Situ Transcription Profiling of Single Cells Reveals Spatial Organization of Cells in the Mouse Hippocampus. Neuron 2016, 92, 342–357. [Google Scholar] [CrossRef] [PubMed]
  7. Eng, C.-H.L.; Lawson, M.; Zhu, Q.; Dries, R.; Koulena, N.; Takei, Y.; Yun, J.; Cronin, C.; Karp, C.; Yuan, G.C.; et al. Transcriptome-scale super-resolved imaging in tissues by RNA seqFISH. Nature 2019, 568, 235–239. [Google Scholar] [CrossRef] [PubMed]
  8. Moffitt, J.R.; Bambah-Mukku, D.; Eichhorn, S.W.; Vaughn, E.; Shekhar, K.; Perez, J.D.; Rubinstein, N.D.; Hao, J.; Regev, A.; Dulac, C.; et al. Molecular, spatial, and functional single-cell profiling of the hypothalamic preoptic region. Science 2018, 362, eaau5324. [Google Scholar] [CrossRef] [PubMed]
  9. Lee, J.H.; Daugharthy, E.R.; Scheiman, J.; Kalhor, R.; Yang, J.L.; Ferrante, T.C.; Terry, R.; Jeanty, S.S.F.; Li, C.; Amamoto, R.; et al. Highly multiplexed subcellular RNA sequencing in situ. Science 2014, 343, 1360–1363. [Google Scholar] [CrossRef]
  10. Moses, L.; Pachter, L. Museum of spatial transcriptomics. Nat. Methods 2022, 19, 534–546. [Google Scholar] [CrossRef]
  11. Stahl, P.L.; Salmen, F.; Vickovic, S.; Lundmark, A.; Navarro, J.F.; Magnusson, J.; Giacomello, S.; Asp, M.; Westholm, J.O.; Huss, M.; et al. Visualization and analysis of gene expression in tissue sections by spatial transcriptomics. Science 2016, 353, 78–82. [Google Scholar] [CrossRef] [PubMed]
  12. Stickels, R.R.; Murray, E.; Kumar, P.; Li, J.; Marshall, J.L.; Di Bella, D.J.; Arlotta, P.; Macosko, E.Z.; Chen, F. Highly sensitive spatial transcriptomics at near-cellular resolution with Slide-seqV2. Nat. Biotechnol. 2021, 39, 313–319. [Google Scholar] [CrossRef] [PubMed]
  13. Vickovic, S.; Eraslan, G.; Salmen, F.; Klughammer, J.; Stenbeck, L.; Schapiro, D.; Aijo, T.; Bonneau, R.; Bergenstrahle, L.; Navarro, J.F.; et al. High-definition spatial transcriptomics for in situ tissue profiling. Nat. Methods 2019, 16, 987–990. [Google Scholar] [CrossRef] [PubMed]
  14. Cho, C.-S.; Xi, J.; Si, Y.; Park, S.-R.; Hsu, J.-E.; Kim, M.; Jun, G.; Kang, H.M.; Lee, J.H. Microscopic examination of spatial transcriptome using Seq-Scope. Cell 2021, 184, 3559–3572.e22. [Google Scholar] [CrossRef] [PubMed]
  15. Chen, A.; Liao, S.; Cheng, M.; Ma, K.; Wu, L.; Lai, Y.; Qiu, X.; Yang, J.; Xu, J.; Hao, S.; et al. Spatiotemporal transcriptomic atlas of mouse organogenesis using DNA nanoball-patterned arrays. Cell 2022, 185, 1777–1792.e21. [Google Scholar] [CrossRef] [PubMed]
  16. Ortiz, C.; Navarro, J.F.; Jurek, A.; Martin, A.; Lundeberg, J.; Meletis, K. Molecular atlas of the adult mouse brain. Sci. Adv. 2020, 6, eabb3446. [Google Scholar] [CrossRef] [PubMed]
  17. Zeisel, A.; Hochgerner, H.; Lonnerberg, P.; Johnsson, A.; Memic, F.; van der Zwan, J.; Haring, M.; Braun, E.; Borm, L.E.; La Manno, G.; et al. Molecular Architecture of the Mouse Nervous System. Cell 2018, 174, 999–1014.e22. [Google Scholar] [CrossRef] [PubMed]
  18. Zhao, E.; Stone, M.R.; Ren, X.; Guenthoer, J.; Smythe, K.S.; Pulliam, T.; Williams, S.R.; Uytingco, C.R.; Taylor, S.E.B.; Nghiem, P.; et al. Spatial transcriptomics at subspot resolution with BayesSpace. Nat. Biotechnol. 2021, 39, 1375–1384. [Google Scholar] [CrossRef] [PubMed]
  19. Li, Z.; Zhou, X. BASS: Multi-scale and multi-sample analysis enables accurate cell type clustering and spatial domain detection in spatial transcriptomic studies. Genome Biol. 2022, 23, 168. [Google Scholar] [CrossRef]
  20. Hu, J.; Li, X.; Coleman, K.; Schroeder, A.; Ma, N.; Irwin, D.J.; Lee, E.B.; Shinohara, R.T.; Li, M. SpaGCN: Integrating gene expression, spatial location and histology to identify spatial domains and spatially variable genes by graph convolutional network. Nat. Methods 2021, 18, 1342–1351. [Google Scholar] [CrossRef]
  21. Dong, K.; Zhang, S. Deciphering spatial domains from spatially resolved transcriptomics with an adaptive graph attention auto-encoder. Nat. Commun. 2022, 13, 1739. [Google Scholar] [CrossRef] [PubMed]
  22. Long, Y.; Ang, K.S.; Li, M.; Chong, K.L.K.; Sethi, R.; Zhong, C.; Xu, H.; Ong, Z.; Sachaphibulkij, K.; Chen, A.; et al. Spatially informed clustering, integration, and deconvolution of spatial transcriptomics with GraphST. Nat. Commun. 2023, 14, 1155. [Google Scholar] [CrossRef] [PubMed]
  23. Ren, H.; Walker, B.L.; Cang, Z.; Nie, Q. Identifying multicellular spatiotemporal organization of cells with SpaceFlow. Nat. Commun. 2022, 13, 4076. [Google Scholar] [CrossRef] [PubMed]
  24. Zong, Y.; Yu, T.; Wang, X.; Wang, Y.; Hu, Z.; Li, Y. conST: An interpretable multi-modal contrastive learning framework for spatial transcriptomics. bioRxiv 2022. [Google Scholar] [CrossRef]
  25. Shang, L.; Zhou, X. Spatially aware dimension reduction for spatial transcriptomics. Nat. Commun. 2022, 13, 7203. [Google Scholar] [CrossRef] [PubMed]
  26. Scrucca, L.; Fop, M.; Murphy, T.B.; Raftery, A.E. mclust 5: Clustering, Classification and Density Estimation Using Gaussian Finite Mixture Models. R J. 2016, 8, 289–317. [Google Scholar] [CrossRef] [PubMed]
  27. McInnes, L.; Healy, J.; Melville, J. UMAP: Uniform Manifold Approximation and Projection for Dimension Reduction. arXiv 2018, arXiv:1802.03426. [Google Scholar] [CrossRef]
  28. Wolf, F.A.; Hamey, F.K.; Plass, M.; Solana, J.; Dahlin, J.S.; Göttgens, B.; Rajewsky, N.; Simon, L.; Theis, F.J. PAGA: Graph abstraction reconciles clustering with trajectory inference through a topology preserving map of single cells. Genome Biol. 2019, 20, 59. [Google Scholar] [CrossRef] [PubMed]
  29. Cao, J.; Spielmann, M.; Qiu, X.; Huang, X.; Ibrahim, D.M.; Hill, A.J.; Zhang, F.; Mundlos, S.; Christiansen, L.; Steemers, F.J.; et al. The single-cell transcriptional landscape of mammalian organogenesis. Nature 2019, 566, 496–502. [Google Scholar] [CrossRef]
  30. Eraslan, G.; Simon, L.M.; Mircea, M.; Mueller, N.S.; Theis, F.J. Single-cell RNA-seq denoising using a deep count autoencoder. Nat. Commun. 2019, 10, 390. [Google Scholar] [CrossRef]
  31. Lopez, R.; Regier, J.; Cole, M.B.; Jordan, M.I.; Yosef, N. Deep generative modeling for single-cell transcriptomics. Nat. Methods 2018, 15, 1053–1058. [Google Scholar] [CrossRef]
  32. Xu, H.; Fu, H.; Long, Y.; Ang, K.S.; Sethi, R.; Chong, K.; Li, M.; Uddamvathanak, R.; Lee, H.K.; Ling, J.; et al. Unsupervised spatially embedded deep representation of spatial transcriptomics. Genome Medicine 2024, 16, 12. [Google Scholar] [CrossRef] [PubMed]
  33. Larsson, L.; Frisen, J.; Lundeberg, J. Spatially resolved transcriptomics adds a new dimension to genomics. Nat. Methods 2021, 18, 15–18. [Google Scholar] [CrossRef] [PubMed]
  34. Munkres, J. Algorithms for the Assignment and Transportation Problems. J. Soc. Ind. Appl. Math. 1957, 5, 32–38. [Google Scholar] [CrossRef]
  35. Korsunsky, I.; Millard, N.; Fan, J.; Slowikowski, K.; Zhang, F.; Wei, K.; Baglaenko, Y.; Brenner, M.; Loh, P.R.; Raychaudhuri, S. Fast, sensitive and accurate integration of single-cell data with Harmony. Nat. Methods 2019, 16, 1289–1296. [Google Scholar] [CrossRef] [PubMed]
  36. Saunders, A.; Macosko, E.Z.; Wysoker, A.; Goldman, M.; Krienen, F.M.; de Rivera, H.; Bien, E.; Baum, M.; Bortolin, L.; Wang, S.; et al. Molecular Diversity and Specializations among the Cells of the Adult Mouse Brain. Cell 2018, 174, 1015–1030.e16. [Google Scholar] [CrossRef] [PubMed]
  37. Beiersdorfer, A.; Wolburg, H.; Grawe, J.; Scheller, A.; Kirchhoff, F.; Lohr, C. Sublamina-specific organization of the blood brain barrier in the mouse olfactory nerve layer. Glia 2020, 68, 631–645. [Google Scholar] [CrossRef] [PubMed]
  38. Van der Maaten, L.; Hinton, G. Visualizing data using t-SNE. J. Mach. Learn. Res. 2008, 9, 2579–2605. [Google Scholar]
  39. Wu, S.Z.; Al-Eryani, G.; Roden, D.L.; Junankar, S.; Harvey, K.; Andersson, A.; Thennavan, A.; Wang, C.; Torpy, J.R.; Bartonicek, N.; et al. A single-cell and spatially resolved atlas of human breast cancers. Nat. Genet. 2021, 53, 1334–1347. [Google Scholar] [CrossRef]
  40. Kleshchevnikov, V.; Shmatko, A.; Dann, E.; Aivazidis, A.; King, H.W.; Li, T.; Elmentaite, R.; Lomakin, A.; Kedlian, V.; Gayoso, A.; et al. Cell2location maps fine-grained cell types in spatial transcriptomics. Nat. Biotechnol. 2022, 40, 661–671. [Google Scholar] [CrossRef]
  41. Hao, Y.; Hao, S.; Andersen-Nissen, E.; Mauck, W.M., 3rd; Zheng, S.; Butler, A.; Lee, M.J.; Wilk, A.J.; Darby, C.; Zager, M.; et al. Integrated analysis of multimodal single-cell data. Cell 2021, 184, 3573–3587.e29. [Google Scholar] [CrossRef] [PubMed]
  42. Bilandzic, M.; Rainczuk, A.; Green, E.; Fairweather, N.; Jobling, T.W.; Plebanski, M.; Stephens, A.N. Keratin-14 (KRT14) Positive Leader Cells Mediate Mesothelial Clearance and Invasion by Ovarian Cancer Cells. Cancers 2019, 11, 1228. [Google Scholar] [CrossRef] [PubMed]
  43. Cheung, K.J.; Padmanaban, V.; Silvestri, V.; Schipper, K.; Cohen, J.D.; Fairchild, A.N.; Gorin, M.A.; Verdone, J.E.; Pienta, K.J.; Bader, J.S.; et al. Polyclonal breast cancer metastases arise from collective dissemination of keratin 14-expressing tumor cell clusters. Proc. Natl. Acad. Sci. USA 2016, 113, E854–E863. [Google Scholar] [CrossRef] [PubMed]
  44. Martinez-Ordonez, A.; Seoane, S.; Avila, L.; Eiro, N.; Macia, M.; Arias, E.; Pereira, F.; Garcia-Caballero, T.; Gomez-Lado, N.; Aguiar, P.; et al. POU1F1 transcription factor induces metabolic reprogramming and breast cancer progression via LDHA regulation. Oncogene 2021, 40, 2725–2740. [Google Scholar] [CrossRef] [PubMed]
  45. Nalio Ramos, R.; Missolo-Koussou, Y.; Gerber-Ferder, Y.; Bromley, C.P.; Bugatti, M.; Nunez, N.G.; Tosello Boari, J.; Richer, W.; Menger, L.; Denizeau, J.; et al. Tissue-resident FOLR2+ macrophages associate with CD8+ T cell infiltration in human breast cancer. Cell 2022, 185, 1189–1207.e25. [Google Scholar] [CrossRef] [PubMed]
  46. Zhang, H.; Wang, Y.; Liu, C.; Li, W.; Zhou, F.; Wang, X.; Zheng, J. The Apolipoprotein C1 is involved in breast cancer progression via EMT and MAPK/JNK pathway. Pathol. Res. Pract. 2022, 229, 153746. [Google Scholar] [CrossRef] [PubMed]
  47. Seo, S.U.; Woo, S.M.; Im, S.-S.; Jang, Y.; Han, E.; Kim, S.H.; Lee, H.; Lee, H.-S.; Nam, J.-O.; Gabrielson, E.; et al. Cathepsin D as a potential therapeutic target to enhance anticancer drug-induced apoptosis via RNF183-mediated destabilization of Bcl-xL in cancer cells. Cell Death Dis. 2022, 13, 115. [Google Scholar] [CrossRef] [PubMed]
  48. Zhang, C.; Zhang, M.; Song, S. Cathepsin D enhances breast cancer invasion and metastasis through promoting hepsin ubiquitin-proteasome degradation. Cancer Lett. 2018, 438, 105–115. [Google Scholar] [CrossRef] [PubMed]
  49. Revel, M.; Sautes-Fridman, C.; Fridman, W.-H.; Roumenina, L.T. C1q+ macrophages: Passengers or drivers of cancer progression. Trends Cancer 2022, 8, 517–526. [Google Scholar] [CrossRef]
  50. Noblejas-Lopez, M.D.M.; Nieto-Jimenez, C.; Morcillo Garcia, S.; Perez-Pena, J.; Nuncia-Cantarero, M.; Andres-Pretel, F.; Galan-Moya, E.M.; Amir, E.; Pandiella, A.; Gyorffy, B.; et al. Expression of MHC class I, HLA-A and HLA-B identifies immune-activated breast tumors with favorable outcome. Oncoimmunology 2019, 8, e1629780. [Google Scholar] [CrossRef]
  51. Nomura, T.; Huang, W.-C.; Zhau, H.E.; Josson, S.; Mimata, H.; WK Chung, L. β2-Microglobulin-mediated signaling as a target for cancer therapy. Anti-Cancer Agents Med. Chem. 2014, 14, 343–352. [Google Scholar] [CrossRef] [PubMed]
  52. Wu, B.; Pan, Y.; Liu, G.; Yang, T.; Jin, Y.; Zhou, F.; Wei, Y. MRPS30-DT Knockdown Inhibits Breast Cancer Progression by Targeting Jab1/Cops5. Front. Oncol. 2019, 9, 1170. [Google Scholar] [CrossRef]
  53. Zeng, B.; Ge, C.; Li, R.; Zhang, Z.; Fu, Q.; Li, Z.; Lin, Z.; Liu, L.; Xue, Y.; Xu, Y.; et al. Knockdown of microsomal glutathione S-transferase 1 inhibits lung adenocarcinoma cell proliferation and induces apoptosis. Biomed. Pharmacother. 2020, 121, 109562. [Google Scholar] [CrossRef] [PubMed]
  54. Huang, X.; Yu, J.; Lai, S.; Li, Z.; Qu, F.; Fu, X.; Li, Q.; Zhong, X.; Zhang, D.; Li, H. Long Non-Coding RNA LINC00052 Targets miR-548p/Notch2/Pyk2 to Modulate Tumor Budding and Metastasis of Human Breast Cancer. Biochem. Genet. 2022, 61, 336–353. [Google Scholar] [CrossRef] [PubMed]
  55. Xiong, D.; Wang, D.; Chen, Y. Role of the long non-coding RNA LINC00052 in tumors. Oncol. Lett. 2021, 21, 316. [Google Scholar] [CrossRef]
  56. Salameh, A.; Fan, X.; Choi, B.K.; Zhang, S.; Zhang, N.; An, Z. HER3 and LINC00052 interplay promotes tumor growth in breast cancer. Oncotarget 2017, 8, 6526–6539. [Google Scholar] [CrossRef]
  57. Grzybowska-Szatkowska, L.; Slaska, B. Mitochondrial NADH dehydrogenase polymorphisms are associated with breast cancer in Poland. J. Appl. Genet. 2014, 55, 173–181. [Google Scholar] [CrossRef]
  58. Lyu, L.; Wang, M.; Zheng, Y.; Tian, T.; Deng, Y.; Xu, P.; Lin, S.; Yang, S.; Zhou, L.; Hao, Q.; et al. Overexpression of FAM234B Predicts Poor Prognosis in Patients with Luminal Breast Cancer. Cancer Manag. Res. 2020, 12, 12457–12471. [Google Scholar] [CrossRef]
  59. Biancalani, T.; Scalia, G.; Buffoni, L.; Avasthi, R.; Lu, Z.; Sanger, A.; Tokcan, N.; Vanderburg, C.R.; Segerstolpe, A.; Zhang, M.; et al. Deep learning and alignment of spatially resolved single-cell transcriptomes with Tangram. Nat. Methods 2021, 18, 1352–1362. [Google Scholar] [CrossRef]
Figure 1. Overview of SECE. (A) SECE is applicable to spatial transcriptome (ST) data with different resolutions, including subcellular, single-cell, near-single-cell and multicellular resolutions. (B) SECE takes gene expression profiles and spatial coordinates as inputs. It begins with an autoencoder (AE) module that compresses gene expression into low-dimensional features based on a count distribution. Then, a graph attention network (GAT) module learns spatial embeddings (SE) by balancing local expression similarity and global spatial proximity, as measured by the distances between expression features and spatial coordinates, respectively. (C) The main SECE outputs include identified spatial domains, low-dimensional visualizations, and inferred spatial trajectories, which are obtained by running clustering, visualization and trajectory inference on the SE.
Figure 1. Overview of SECE. (A) SECE is applicable to spatial transcriptome (ST) data with different resolutions, including subcellular, single-cell, near-single-cell and multicellular resolutions. (B) SECE takes gene expression profiles and spatial coordinates as inputs. It begins with an autoencoder (AE) module that compresses gene expression into low-dimensional features based on a count distribution. Then, a graph attention network (GAT) module learns spatial embeddings (SE) by balancing local expression similarity and global spatial proximity, as measured by the distances between expression features and spatial coordinates, respectively. (C) The main SECE outputs include identified spatial domains, low-dimensional visualizations, and inferred spatial trajectories, which are obtained by running clustering, visualization and trajectory inference on the SE.
Biomolecules 14 00674 g001
Figure 2. Application of SECE to mouse visual cortex STARmap data. (A) Layer structure of the tissue section from the original study. (B) Spatial domains identified by SECE. (C) Spatial domains identified by BASS, SpaceFlow, GraphST, STAGATE, SpatialPCA, SpaGCN, BayesSpace and conST. (D) Assessment of spatial domain identification and SE learning across methods using ARI, NMI and ASW. (E) UMAP visualizations generated by SECE, SpaceFlow, GraphST, STAGATE and SpatialPCA, colored by annotated layers. (F) PAGA graphs generated by SECE, SpaceFlow, GraphST, STAGATE and SpatialPCA.
Figure 2. Application of SECE to mouse visual cortex STARmap data. (A) Layer structure of the tissue section from the original study. (B) Spatial domains identified by SECE. (C) Spatial domains identified by BASS, SpaceFlow, GraphST, STAGATE, SpatialPCA, SpaGCN, BayesSpace and conST. (D) Assessment of spatial domain identification and SE learning across methods using ARI, NMI and ASW. (E) UMAP visualizations generated by SECE, SpaceFlow, GraphST, STAGATE and SpatialPCA, colored by annotated layers. (F) PAGA graphs generated by SECE, SpaceFlow, GraphST, STAGATE and SpatialPCA.
Biomolecules 14 00674 g002
Figure 3. Application of SECE to mouse hippocampus Slide-seqV2 data. (A) Annotation of hippocampus structures from the Allen Brain Atlas (ABA) for adult mouse brain. (B) Left: spatial visualization of domains identified by SECE. Right: UMAP visualization of domains identified by SECE. Spatial domains were annotated based on the ABA structures. (CA, Ammon’s horn; DG, Dentate gyrus.) (C) Spatial visualization of CA1, CA2 and CA3 domains identified by SECE (Left) and the corresponding marker genes Wsf1, Rgs14 and Nptxr (right). (D) The 14 spatial regions identified by BASS, SpaceFlow, GraphST, STAGATE, SpatialPCA, SpaGCN, BayesSpace and conST. (E) Pseudo-time of each cell, calculated by Monocle3 based on SECE, SpaceFlow, GraphST, STAGATE and SpatialPCA embeddings. (F) Pseudo-time of the cells in each isocortex layer based on SECE, SpaceFlow, GraphST, STAGATE and SpatialPCA.
Figure 3. Application of SECE to mouse hippocampus Slide-seqV2 data. (A) Annotation of hippocampus structures from the Allen Brain Atlas (ABA) for adult mouse brain. (B) Left: spatial visualization of domains identified by SECE. Right: UMAP visualization of domains identified by SECE. Spatial domains were annotated based on the ABA structures. (CA, Ammon’s horn; DG, Dentate gyrus.) (C) Spatial visualization of CA1, CA2 and CA3 domains identified by SECE (Left) and the corresponding marker genes Wsf1, Rgs14 and Nptxr (right). (D) The 14 spatial regions identified by BASS, SpaceFlow, GraphST, STAGATE, SpatialPCA, SpaGCN, BayesSpace and conST. (E) Pseudo-time of each cell, calculated by Monocle3 based on SECE, SpaceFlow, GraphST, STAGATE and SpatialPCA embeddings. (F) Pseudo-time of the cells in each isocortex layer based on SECE, SpaceFlow, GraphST, STAGATE and SpatialPCA.
Biomolecules 14 00674 g003
Figure 4. Application of SECE to mouse olfactory bulb Stereo-seq data. (A) Left: annotation of mouse olfactory bulb structures from the ABA. Right: spatial visualization of domains identified by SECE. (RMS, Rostral migratory stream; GCL, Granule cell layer; IPL, Inner plexiform layer; MCL, Mitral cell layer; EPL, External plexiform layer; GL, Glomerular layer; ONL, Olfactory nerve layer.) (B) Heatmap of known marker gene expression for each layer; the median values of the centered and standardized gene expression for each region are shown. The green boxes are the subregions of GCL and ONL, respectively. (C) The 9 spatial regions identified by BASS, SpaceFlow, GraphST, STAGATE, SpatialPCA, SpaGCN, BayesSpace and conST. (D) UMAP visualizations generated by SECE, SpaceFlow, GraphST, STAGATE and SpatialPCA, colored by annotated layers. (E) PAGA graphs generated by SECE, SpaceFlow, GraphST, STAGATE and SpatialPCA.
Figure 4. Application of SECE to mouse olfactory bulb Stereo-seq data. (A) Left: annotation of mouse olfactory bulb structures from the ABA. Right: spatial visualization of domains identified by SECE. (RMS, Rostral migratory stream; GCL, Granule cell layer; IPL, Inner plexiform layer; MCL, Mitral cell layer; EPL, External plexiform layer; GL, Glomerular layer; ONL, Olfactory nerve layer.) (B) Heatmap of known marker gene expression for each layer; the median values of the centered and standardized gene expression for each region are shown. The green boxes are the subregions of GCL and ONL, respectively. (C) The 9 spatial regions identified by BASS, SpaceFlow, GraphST, STAGATE, SpatialPCA, SpaGCN, BayesSpace and conST. (D) UMAP visualizations generated by SECE, SpaceFlow, GraphST, STAGATE and SpatialPCA, colored by annotated layers. (E) PAGA graphs generated by SECE, SpaceFlow, GraphST, STAGATE and SpatialPCA.
Biomolecules 14 00674 g004
Figure 5. Application of SECE to mouse brain Stereo-seq data. (A) The annotation of mouse hemibrain structures from the ABA. (B) Left: spatial visualization of domains identified by SECE. Right: UMAP visualization of domains identified by SECE. Spatial domains were annotated based on ABA. (sl/r, stratum lacunosum/raditum cornu ammonis; DG, dentate gyrus; FT, fiber tract; MLDG, molecular layer of dentate gyrus; MRN, midbrain reticular nucleus; SN, substantia nigra; VTA, ventral tegmental area.) (C) The 20 spatial regions identified by BASS, SpaceFlow, GraphST, STAGATE, SpatialPCA, SpaGCN, BayesSpace and conST.
Figure 5. Application of SECE to mouse brain Stereo-seq data. (A) The annotation of mouse hemibrain structures from the ABA. (B) Left: spatial visualization of domains identified by SECE. Right: UMAP visualization of domains identified by SECE. Spatial domains were annotated based on ABA. (sl/r, stratum lacunosum/raditum cornu ammonis; DG, dentate gyrus; FT, fiber tract; MLDG, molecular layer of dentate gyrus; MRN, midbrain reticular nucleus; SN, substantia nigra; VTA, ventral tegmental area.) (C) The 20 spatial regions identified by BASS, SpaceFlow, GraphST, STAGATE, SpatialPCA, SpaGCN, BayesSpace and conST.
Biomolecules 14 00674 g005
Figure 6. Application of SECE to human breast cancer Visium data. (A) Pathology annotation of the tissue section from the original study. (B) Spatial regions identified by SECE. (C) Fine-grained regions identified by SECE, that is, DCIS/LCIS-inner, DCIS/LCIS-outer, DCIS/LCIS-edge, IDC-inner, IDC-outer and IDC-edge. (D) Cell type compositions of 6 fine-grained regions. (E) Heatmaps of normalized expression of signature genes identified in the DE analysis based on the 6 fine regions. (F) Pseudo-times of spots calculated by Monocle3 in the DCIS/LCIS (top) and IDC regions (bottom).
Figure 6. Application of SECE to human breast cancer Visium data. (A) Pathology annotation of the tissue section from the original study. (B) Spatial regions identified by SECE. (C) Fine-grained regions identified by SECE, that is, DCIS/LCIS-inner, DCIS/LCIS-outer, DCIS/LCIS-edge, IDC-inner, IDC-outer and IDC-edge. (D) Cell type compositions of 6 fine-grained regions. (E) Heatmaps of normalized expression of signature genes identified in the DE analysis based on the 6 fine regions. (F) Pseudo-times of spots calculated by Monocle3 in the DCIS/LCIS (top) and IDC regions (bottom).
Biomolecules 14 00674 g006
Figure 7. Application of SECE to DLPFC Visium data. (A) Pathology annotation of section 151674 from the original study. (B) Spatial regions of the section 151674 identified by SECE. (C) Spatial regions of the section 151674 identified by eight other methods. (D) Boxplot of ARI values for 12 slices (p = 0.02, one-tailed paired t test). * represents the p value less than 0.05. In the boxplot, the center line denotes the median, box limits denote the upper and lower quartiles, and whiskers denote the 1.5× interquartile range. The blue horizontal line indicates the median ARI value of SECE.
Figure 7. Application of SECE to DLPFC Visium data. (A) Pathology annotation of section 151674 from the original study. (B) Spatial regions of the section 151674 identified by SECE. (C) Spatial regions of the section 151674 identified by eight other methods. (D) Boxplot of ARI values for 12 slices (p = 0.02, one-tailed paired t test). * represents the p value less than 0.05. In the boxplot, the center line denotes the median, box limits denote the upper and lower quartiles, and whiskers denote the 1.5× interquartile range. The blue horizontal line indicates the median ARI value of SECE.
Biomolecules 14 00674 g007
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Yu, Y.; He, Y.; Xie, Z. Accurate Identification of Spatial Domain by Incorporating Global Spatial Proximity and Local Expression Proximity. Biomolecules 2024, 14, 674. https://doi.org/10.3390/biom14060674

AMA Style

Yu Y, He Y, Xie Z. Accurate Identification of Spatial Domain by Incorporating Global Spatial Proximity and Local Expression Proximity. Biomolecules. 2024; 14(6):674. https://doi.org/10.3390/biom14060674

Chicago/Turabian Style

Yu, Yuanyuan, Yao He, and Zhi Xie. 2024. "Accurate Identification of Spatial Domain by Incorporating Global Spatial Proximity and Local Expression Proximity" Biomolecules 14, no. 6: 674. https://doi.org/10.3390/biom14060674

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop