Introduction

Biological sex is a major determinant of immune system function and responsiveness under physiological and pathological conditions. Females mount more vigorous innate and adaptive immune responses as compared to males1. As a consequence, males are more vulnerable to infections2, while females are predisposed to autoimmunity3. Sexual dimorphism pertains also to the development and progression of malignancies4, responsiveness to immunizations5 and transplantation6.

The molecular basis of sex disparities in immunity remains incompletely understood, yet it is partially attributed to escapee genes from X chromosome inactivation (XCI) and sex hormones7. Mammalian X chromosomes, for which males are hemizygous, are enriched for immune-related genes8. Some of these genes, such as toll-like receptor (TLR) 7, TLR8, CD40 ligand (CD40LG), interleukin 1 receptor associated kinase 1 (IRAK1), chromosome X open reading frame 21 (CXorf21) and X-linked inhibitor of apoptosis (XIAP), escape the XCI process, thus disrupting dosage compensation and resulting in higher expression levels in females9,10,11. On the other hand, estrogens can modulate Toll-like receptor (TLR) responses and the synthesis of certain immune genes9,12,13,14.

Biological sex and sex-chromosome dosage exert widespread effects on the transcriptome of blood immune cells15,16,17,18,19,20,21,22,23, which may be partially explained by sex-specific epigenetic modifications and genome regulation1,9,24. This has been suggested by the in-silico analysis of transcription factor binding sites and inference of regulatory networks that control sex-biased gene expression in immune cells and other tissues15,21,22. However, there is currently limited evidence directly linking genome regulatory proteins to the shaping of cell-specific molecular programs, particularly in relation to sex-biased immune phenotypes25,26.

Systemic Lupus Erythematosus (SLE) is an archetype sex-biased autoimmune disease, with a female-to-male ratio as high as 10:127,28. Augmented type I interferon (IFN) and TLR signaling drive the activation of lupus myeloid cells, contributing to the breakdown of self-tolerance through the maturation and expansion of autoreactive lymphocytes29,30. Work by us and others has positioned monocytes centrally to SLE pathogenesis, as they display marked molecular and functional aberrations31,32,33,34, secrete inflammatory cytokines35,36,37 and infiltrate target tissues38,39. Despite advances, patients with SLE face excessive morbidity and poor long-term outcomes, thus underlining the urgent need to expand the disease therapeutic armamentarium40.

Previously, we probed the blood transcriptome in SLE and healthy individuals, aiming to explore the molecular regulation of the disease41. Through differential gene expression analysis between sexes, we observed that SMC1A exhibited the most prominent sex-biased expression in SLE compared to healthy counterparts. SMC1A encodes for the nuclear protein “Structural Maintenance of Chromosomes 1A” and resides on the Xp11.22 region that partially escapes XCI42,43. Together with other subunits (SMC3, RAD21, STAG1/2), SMC1A forms the cohesin complex involved in chromosome segregation, homologous recombination, and genome regulation44,45,46. Cohesin mediates chromatin loop formation to support enhancer-promoter communication and coordinate gene transcription50. It also interacts with transcription factors and facilitates their binding to DNA46,47,48,49,50, a fact suggestive of a putative direct regulatory role.

The implication of SMC1A/cohesin in human pathophysiology is highlighted by the discovery of mutations in these genes, which are associated with various malignancies and developmental disorders, collectively known as cohesinopathies51. Despite some uncertainties, these pathologies are likely driven by perturbed gene expression due to altered function or diminished expression of cohesin52,53,54,55. Notably, cohesin mediates TLR-inducible inflammatory molecular programs in myeloid cells56,57 and binds to the IFNG locus to regulate its expression in CD4+ T-cells58. Nonetheless, it remains elusive whether SMC1A/cohesin plays a role in more generalized transcriptional and functional deregulations of immune cells involved in autoimmunity.

Here, we aim to delineate the role of SMC1A in autoimmune-related inflammation and the effect of female-biased SMC1A in SLE pathogenesis. Moreover, we seek to gain insights into the broader mechanisms that may underlie the escape of SMC1A from XCI, a process that remains poorly defined. We show that, in response to lupus-relevant stimuli, SMC1A binds preferentially to active enhancers of immune and inflammatory genes in monocytes and regulates cytokine production, including IL6 and IL1A. Silencing of SMC1A downregulates lupus-related genes, while transcriptome profiling of SLE patient monocytes reveals female-prominent activation of SMC1A targets. These data support a key role for SMC1A as a sex-biased chromatin modifier that shapes the inflammatory gene program of monocytes and contributes to susceptibility to SLE and other autoimmune disorders.

Results

Monocytes from SLE patients exhibit enhanced female-biased expression of SMC1A

In a previous unbiased blood transcriptome analysis41, we showed that SMC1A encoding for Structural Maintenance of Chromosomes 1A, displayed strong female-biased expression in SLE. To corroborate this finding, we combined our original cohort with blood RNA-sequencing results from additional SLE and healthy individuals. We reached a total of 15 males with SLE who were compared with an equal number of female SLE, healthy female and male counterparts thus forming a balanced dataset (Supplementary Table 1) (GEO accession number GSE279728). Following batch-effect correction, differential gene expression analysis revealed significant female-biased SMC1A mRNA levels in SLE (log2-fold change [log2FC] = 0.39, p = 0.0131) compared to healthy state (log2FC = 0.13, p = 0.2059) (Fig. 1A), thus prompting us to investigate the role of SMC1A in lupus.

Fig. 1: SMC1A displays augmented female-biased expression in lupus monocytes.
figure 1

A SMC1A levels (DESeq2 normalized counts) in blood RNA-sequencing from male and female SLE and healthy individuals (n = 15/group). Log2-fold change for the male versus female comparison, with corresponding p-values, were generated using DESeq2 without adjustment for multiple comparisons. B TaqMan Assays was performed to assay differential SMC1A mRNA levels in blood monocytes from clusters of healthy (n = 10 males, 8 females) and SLE individuals (n = 18 males, 21 females). Mixed model analysis accounted for nested data structure and inter-individual variability, introducing sex (male versus female), state (SLE versus healthy) and the sex × state interaction as within-cluster covariates. Least Squares Means method was used for post-hoc sex comparisons without multiple-testing adjustment. C Western immunoblotting for SMC1A and GAPDH in blood monocytes from SLE patients (n = 4 male/female pairs). The lower lane provides a summary of the SMC1A levels were normalized by GAPDH. Mixed model analysis was performed as above. D SMC1A expression levels (normalized) in the blood from male (n = 27) and female (n = 25) patients with ankylosing spondylitis (AS). The GEO dataset GSE73754 was analyzed using shinyGEO (ILMN_1652006 probe). P-value was produced from unpaired Student’s t test. E RT-qPCR (SYBR Green) was performed to assay SMC1A mRNA levels in blood monocytes from AS patients (n = 16 male/female pairs) and healthy individuals (n = 9 male/female pairs). Mixed model analysis was performed as in (B). F SMC1A levels (normalized) in blood RNA-sequencing from 46,XX females (n = 15), males with Klinefelter syndrome (47,XXY) (n = 35) and 46,XY males (n = 15). Dataset GSE42331 was analyzed and expression levels were normalized with the Bioinformatics Array Research Tool. ANOVA was performed with post-hoc Tukey’s test. G Experimental setup to generate SLE-like monocytes. Healthy CD14+ monocytes were pre-treated with IFNα and TNF for 18 h, followed by LPS for 1–4 h. Created in BioRender. H SMC1A mRNA (RT-qPCR) in blood monocytes from healthy individuals (n = 13 male/female pairs), either untreated (basal state) or stimulated as described in (G) (SLE-like). Mixed model analysis was performed as described in (B). In all panels, dots represent different individuals, and bar plots show the mean ± standard error (SE) expression. All p-values are two-tailed.

To explore whether sex divergence in SMC1A is driven by one or more circulating immune cell types, blood was collected from pairs of male and female SLE patients with low disease activity (Supplementary Table 2), as well as from age-matched healthy counterparts. We separated neutrophils, CD14+ monocytes, CD4+ T cells and CD19+ B cells, which constitute the majority of white blood cells and represent major subsets implicated in lupus, followed SMC1A mRNA quantification by quantitative RT-PCR (RT-qPCR). Mixed-effects analysis was used to account for nested data structure and inter-individual variability. Both T- and B-lymphocytes exhibited female overexpression of SMC1A, which, however, was comparable between the SLE and healthy state (Supplementary Fig. 1A). In contrast, CD14+ monocytes showed greater female-male difference in SMC1A mRNA levels in SLE compared to healthy counterparts, as indicated by a significant sex × disease state interaction term (β = 0.184, p = 0.026) (Fig. 1B, Supplementary Table 3). This exaggerated female-biased SMC1A expression in SLE compared to healthy monocytes mirrored the whole blood data (Fig. 1A). A similar pattern of enhanced female-biased SMC1A mRNA levels was also observed in blood CD14+ monocytes (GSE137143) from patients with multiple sclerosis—another autoimmune disorder displaying female susceptibility59—compared to those from healthy individuals (Supplementary Fig. 1B). SLE, especially active disease, had an overall negative effect on SMC1A expression (Supplementary Tables 34, Supplementary Fig. 1C). Western immunoblotting in SLE patient-derived blood CD14+ monocytes confirmed the increased expression of SMC1A in females (Fig. 1C). These findings, coupled with evidence that transcriptional sexual dimorphism is particularly evident in monocytes/macrophages and increases after stimulation with interferon20,60—a key mediator of SLE61—guided our decision to focus on monocytes.

To address the specificity of our observation, we used ankylosing spondylitis (AS) as a control autoimmune disease displaying modest male predominance62,63. First, we analyzed publicly available blood transcriptome data from AS patients (GSE73754) and found no sex variation in SMC1A (Fig. 1D). Next, we recruited male and female AS patients and healthy individuals (Supplementary Table 5) and purified blood CD14+ monocytes. RT-qPCR revealed no difference in SMC1A mRNA levels depending on patient sex (Fig. 1E), therefore suggesting that female-biased expression of SMC1A is accentuated in the monocytes of the female-predominant autoimmune disease of lupus, but in AS.

Sex-biased SMC1A expression in SLE is determined by cell-intrinsic and -extrinsic factors

SMC1A partially escapes XCI resulting in elevated basal gene expression in female cells17,20,64,65. This agrees with our RT-qPCR data showing increased SMC1A mRNA levels in healthy females than males across all white blood cell subsets except for neutrophils (Fig. 1B). To further implicate an X-chromosome effect on SMC1A, we explored blood transcriptomic data (GSE42331) from males with Klinefelter syndrome (KS; 47,XXY), an aneuploid genetic condition characterized by an additional X chromosome copy and increased tendency for SLE66,67. SMC1A was significantly elevated in KS as compared with 46,XY males, and comparable to 46,XX female counterparts (Fig. 1F).

Next, we assessed whether sex-biased SMC1A expression may be affected by exogenous factors pertinent to lupus. First, we mimicked the disease by culturing CD14+ monocytes from healthy, age-matched women and men with sequential IFN-α/TNF and LPS challenge (Fig. 1G) following a previously described protocol68 with slight modifications. This set-up reproduced the inflammatory environment of lupus as evidenced by the induction of the hallmark cytokines IL-1b, IL-6 and CXCL10 and the membrane markers CD80 and HLA-DR (Supplementary Fig. 2). Compared to unstimulated condition, lupus-inducing stimulation reduced SMC1A in both sexes (Fig. 1H), however, the female-biased expression of SMC1A persisted in the lupus-like state, thus resembling the pattern observed in SLE patient monocytes (Fig. 1B). We also examined whether SMC1A expression may be affected by hormonal factors such as estrogen, which can modulate the expression of inflammation-related genes, thus conferring risk to autoimmunity12,13,69,70. Addition of estradiol to blood CD14+ monocytes from healthy donors treated with healthy or SLE patient-derived serum had no significant effect on SMC1A mRNA levels (Supplementary Fig. 3). Collectively, these findings suggest that in addition to X-chromosome gene-dosage, inflammatory cues pertinent to SLE might contribute to the female-biased SMC1A expression observed in the disease.

SMC1A dosage affects the expression of immune response genes in monocytes under basal and lupus-like conditions

Research involving genetic deletion or silencing of individual cohesin subunits has unveiled their role in a number of biological processes including regulation of inducible gene expression56,57. To address the functional consequences of SMC1A levels in our disease context, we performed RNA silencing by nucleofection in basal (unstimulated) and lupus-like monocytes, followed by whole transcriptome profiling (Fig. 2A) (GEO accession number GSE279728). Experiments were performed on blood CD14+ monocytes from three independent healthy males, used as the reference sex since SMC1A expression was most affected in males with SLE. Silencing efficiency was assessed by RT-qPCR and western blot analysis (we detected an average of 56% reduction in SMC1A protein levels), and cell viability by Annexin-7AAD staining (Fig. 2B, Supplementary Fig. 4).

Fig. 2: SMC1A dosage affects the expression of immune-related genes.
figure 2

A CD14+ monocytes purified from healthy individuals were transfected with siRNA against SMC1A for 4 h, then pre-treated with recombinant IFNα and TNF for 18 h, followed by a 3-h LPS challenge. Created in BioRender. B RT-qPCR was used to measure SMC1A mRNA levels in SMC1A-silenced (SMC1A si) or control-silenced (ctrl si; scramble) CD14+ monocytes from healthy individuals (n = 3), which were either untreated (basal state) or stimulated (SLE-like state). Quantification was performed using the 2–ΔΔCT method (ΔCt = SMC1A Ct minus 18S Ct). Each dot represents a different donor, and bar plots show the mean ± standard error (SE) expression. ANOVA was performed followed by post-hoc Tukey’s test. C Volcano plot of DEGs from RNA sequencing analysis of SMC1A-silenced (SMC1A si) compared to control-silenced (ctrl si; scramble) CD14+ monocytes in the basal state (triplicates, p-value < 0.05). Log2 fold change (log2FC) along with the corresponding p-values were generated using DESeq2. The y-axis represents the –log10(p-values), while the x-axis shows the log2FC in expression levels of genes in SMC1A-silenced versus control-silenced CD14+ monocytes. D Volcano plot of DEGs from RNA sequencing analysis of SMC1A-silenced (SMC1A si) compared to control-silenced (ctrl si; scramble) CD14+ monocytes in the SLE-like state (triplicates, p < 0.05). The y-axis represents the –log10(p-values), while the x-axis shows the log2FC in expression levels of genes in SMC1A-silenced versus control-silenced CD14+ monocytes (generated using DESeq2). E Scatter plot of gene expression changes (Log2FC; SMC1A-silenced versus control-silenced cells) in CD14+ monocytes under basal (y-axis) and SLE-like (x-axis) conditions. Analysis was performed in the combined list of 860 DEGs shown in (C, D). Green dots represent DEGs specific to basal-state monocytes, blue to SLE-like monocytes, and yellow to both. F Enrichment analysis for over-represented Gene Ontology (GO) terms in DEGs upon SMC1A silencing in basal and SLE-like monocytes. All p-values are two-tailed.

By comparing the transcriptome profiles of SMC1A-silenced versus control monocytes under basal and lupus-like conditions, we identified 442 and 418 differentially expressed genes (DEG) (p < 0.05) respectively, with 34 shared genes between the two states (Fig. 2C, D). In these DEGs, SMC1A downregulation led to overall modest alterations in the mRNA levels (log2FC values represented on the x-axis of Fig. 2C, D plots), consistent with previous studies examining transcriptional changes following elimination of other cohesin members57,71. The 826 non-shared DEGs displayed uncorrelated expression patterns between the two states (Fig. 2E; compare the blue- and green-colored dots), suggesting that SMC1A acquires distinct transcriptional associations during transition from basal to inflammatory condition. To gain functional insights, we performed pathway enrichment analysis of the SMC1A-related DEGs. We found a strong prevalence of immune-related pathways, including regulation of the MAPK cascade, response to type II interferon, immune response-regulating signaling pathway, cell chemotaxis/migration and cellular response to biotic stimulus (Fig. 2F).

To further explore the relationship between SMC1A and lupus, we focused on genes that are differentially expressed in lupus-like versus unstimulated monocytes (|log2FC| ≥ 1, adjusted p < 0.05) (GEO accession number GSE279728). These 2778 DEGs, which we term lupus-responsive genes (Supplementary Data 1), showed strong transcriptional concordance with DEGs identified in lupus-like monocytes from a previous study68 (Supplementary Fig. 5) and in CD14⁺ monocytes from SLE patients compared to healthy controls (Fig. 3A). Moreover, lupus-responsive genes were enriched for similar hallmark inflammatory signatures as those observed in SLE patient-derived monocytes, further corroborating the disease relevance of our ex vivo lupus-like system (Fig. 3B). SMC1A-related DEGs overlapped significantly with lupus-responsive genes in both basal (55% overlap; odds ratio 4.4, p = 1.6e-69, Fisher’s exact test) and lupus-like (37% overlap; odds ratio 2.0, p = 4.2e-12) conditions. Furthermore, by summarizing the expression level changes of lupus-responsive genes, we observed that genes repressed by SMC1A downregulation tend to become transcriptionally activated in lupus-like monocytes, and vice versa (Fig. 3C, compare the fourth with the third column, and the second with the first column). Together, these data suggest that SMC1A may impact immune response genes in monocytes and that genes deregulated during lupus are particularly susceptible to SMC1A gene levels.

Fig. 3: Lupus-responsive genes are sensitive to SMC1A levels.
figure 3

A Comparison of lupus-responsive genes (significantly altered in SLE-like vs. basal CD14⁺ monocytes; triplicates, |log2FC| ≥ 1, adjusted p < 0.05) with DEGs from RNA-sequencing of CD14⁺ monocytes from SLE patients (n = 10, unselected for sex) versus healthy controls (n = 9, unselected for sex) using the same thresholds (|log2FC| ≥ 1, adjusted p < 0.05). The y-axis shows log2FC in SLE-like versus basal monocytes, and the x-axis shows log2FC in SLE versus healthy monocytes. Each dot represents a gene. Genes commonly downregulated fall in the lower-left quadrant, while commonly upregulated genes fall in the upper-right quadrant. Fisher’s exact test revealed significant overlap between transcriptionally activated lupus-responsive genes and SLE-specific DEGs (p = 2.0e-148). B Functional enrichment analysis (Gene Ontology terms) of lupus-responsive genes. The left panel shows enriched pathways among DEGs in SLE versus healthy CD14⁺ monocytes; the right panel shows enriched pathways among lupus-responsive genes (SLE-like versus basal monocytes). C Dot plot showing log2FC changes in lupus-responsive genes that were also differentially expressed upon SMC1A silencing versus control silencing in CD14⁺ monocytes (Fig. 2C, D). Results are presented separately for genes up- or down-regulated in SMC1A-silenced versus control-silenced cells under basal (left two panels) or SLE-like (right two panels) conditions. Horizontal lines indicate median log2FC values. One-way ANOVA was performed, followed by post-hoc comparisons using the Least Squares Means method. All p-values are two-tailed.

SMC1A is redistributed to active distal regulatory elements during lupus inflammation

SMC1A is part of the cohesin complex that regulates genome structure and expression in a cell- and context-specific manner50. To obtain a reference map of SMC1A targets across the genome, we isolated blood CD14+ monocytes from independent healthy males (used as a reference sex) and performed chromatin immunoprecipitation followed by DNA sequencing (ChIP-seq) under basal and lupus-like conditions (Supplementary Fig. 6A–C) (GEO accession number GSE279726). Cells were assayed at 1-h post-LPS challenge to study the effect of SMC1A/cohesin binding on the immediate responsive genes. Such brief activation of TLR-4 by LPS may induce a gene transcription program but not proliferation61,67, thus uncoupling the regulatory role of SMC1A from its function during sister chromatid cohesion.

Differential binding analysis in lupus-like versus unstimulated monocytes (|log2FC| > 1, p < 0.001), revealed significant enrichment of SMC1A binding at 6,839 genomic regions and a corresponding depletion at 1131 regions, suggestive of a disease-specific rewiring of SMC1A interactions with chromatin (Fig. 4A). Annotation of these regions showed that ~45% of the peaks with increased SMC1A signal in lupus-like monocytes resided within distal intergenic regions (>1 kb from the transcription start site [TSS]), whereas 75% of the SMC1A peaks with decreased signal were at gene promoters (<1 kb from the TSS) (Fig. 4B).

Fig. 4: The chromatin occupancy of SMC1A is significantly altered in lupus inflammation.
figure 4

A Volcano plot of differential genomic binding events (|log2FC| > 1, p < 0.001 generated using DESeq2) from SMC1A ChIP-sequencing analysis in SLE-like (n = 3) versus basal (n = 2) male CD14+ monocytes. Blue dots represent regions with reduced SMC1A binding (n = 1131) and red dots with increased SMC1A binding (n = 6839). B Stacked bar graph showing the genomic distribution of reduced (upper lane) and increased (lower lane) SMC1A binding events in SLE-like versus basal monocytes from the ChIP-seq analysis shown in (A). C Proportion of genes associated with promoter and enhancer genomic regions that demonstrate reduced (down-bound) or increased (up-bound) SMC1A binding during transition from basal to SLE-like monocytes. D Volcano plot illustrating H3K27ac signal changes (|log2FC| > 1, p < 0.001 generated using DESeq2) from ChIP-sequencing analysis in SLE-like (n = 4) versus basal (n = 3) CD14+ monocytes. Blue dots represent regions with reduced (n = 274), and red dots with increased (n = 502) H3K27ac signal. E Area-proportional euler diagram displaying the intersection of genes associated with differential—increased (designated with up-arrow) or reduced (designated with down-arrow)—SMC1A binding and H3K27ac signal. The majority (295 out of 440, 67%) of genes with increased H3K27ac signal showed enriched SMC1A binding. F Functional enrichment analysis (Gene Ontology terms) of genes associated with both increased SMC1A binding and increased H3K27ac signal in SLE-like compared to basal monocytes. All p-values are two-tailed.

To integrate these results into the context of the monocyte transcriptome, we assessed the enrichment/depletion of SMC1A binding at the regulatory elements of an extensive list of genes (n = 11,971), for which we had gene expression measurements from our RNA-sequencing assays. By utilizing the FANTOM572 and GeneHancer72 databases to annotate and link enhancer elements to genes, we found that over 7% of genes (n = 858; Supplementary Fig. 6D) exhibited increased SMC1A binding at their enhancers versus 2.4% (n = 285), which showed a decreased signal upon stimulation. By contrast, only 1% of the genes (n = 125) showed increased SMC1A binding and 4.9% (n = 585 genes; Supplementary Fig. 6E) displayed depleted SMC1A signal at their promoters (Fig. 4C). Thus, under lupus inflammation, there is a prominent genomic redistribution of SMC1A characterized by increased binding at enhancer elements and reciprocal depletion from gene promoters.

To further explore the regulatory dynamics associated with SMC1A binding, we profiled the genome-wide H3K27ac (histone 3 lysine 27 acetylation) marks, a well-characterized marker for active chromatin68, by ChIP-sequencing in CD14+ monocytes from independent male donors cultured under basal and lupus-like conditions (Supplementary Fig. 6F) (GEO accession number GSE279726). The latter exhibited extensive changes, with increased H3K27ac signal in 502 genomic regions (log2FC > 1, p < 0.001) (Fig. 4D), consistent with previously described epigenome perturbations in lupus disease69. These activated regions involved a total of 440 genes with functional enrichment for immune pathways such as regulation of T-cell activation and response to virus (Supplementary Fig. 6G).

Combining the SMC1A and H3K27ac ChIP-seq datasets, we identified a significant enrichment of SMC1A up-bound genes within the H3K27ac-enriched gene set (p = 1.2e-14, odds ratio = 6.16) (Fig. 4E). This result suggests a notable association between SMC1A binding and H3K27ac activity at these loci. Furthermore, the 295 genes with co-enriched SMC1A and H3K27ac signals were enriched in pathways related to cytokine signaling and response to virus (Fig. 4F). These data suggest a possible role for SMC1A in regulating distal enhancer activity linked to the immune-related gene program in monocytes during lupus inflammation.

The cohesin complex can facilitate and maintain chromatin remodeling and accessibility57,73,74,75. Likewise, the chromatin environment might impact the binding and functional consequences of cohesin76,77. To this end, we analyzed publicly available chromatin accessibility features assayed by transposase-accessible chromatin with sequencing (ATAC-seq) in the same experimental setup of lupus-like monocytes68. Integrative analysis was performed to construct a comprehensive genomic map of ATAC-seq and H3K27ac peaks indicative of open and active chromatin regions, respectively (Fig. 5A). These data, combined with the differential binding events of SMC1A, confirmed the accentuated redistribution of SMC1A within distal active regulatory elements during transition from basal to lupus-like state (Fig. 5B). Specifically, more than 8% of genes with active enhancers/promoters (based on H3K27ac signal) (n = 9834) displayed enriched SMC1A binding at their enhancers (n = 851; odds ratio 29.2, p < 2.2e-16, by Fisher’s exact test) and more than 5% showed decreased SMC1A binding at their promoters (n = 559; odds ratio 4.97, p < 2.2e-16, by Fisher’s exact test). Therefore, the binding dynamics of SMC1A are strongly correlated with the activity and accessibility of the underlying regulatory element.

Fig. 5: During lupus inflammation SMC1A is redistributed to active enhancers.
figure 5

A Schematic representation of the integration analysis of ATAC-sequencing (GEO dataset GSE100383; n = 3 lupus-like monocytes), SMC1A (Fig. 4A) and H3K27ac (Fig. 4C) ChIP-sequencing data in SLE-like and unstimulated (basal state) CD14+ monocytes. Chromatin activity states combined with SMC1A binding enrichment at specific genome regions were used to define the putative regulatory role of SMC1A in the context of lupus. Created in BioRender. B Genome-wide map of ATAC-seq, H3K27ac ChIP-seq and SMC1A ChIP-seq peaks in monocytes cultured under SLE-like and basal conditions. Differentially SMC1A-bound regions were intersected with Ensembl Transcription Start Sites (TSS) and FANTOM572 enhancer elements (enhancer TSSs [eTSSs]), each extended by 500 base pairs, upstream and downstream. Only promoters from transcribed genes were kept in the dataset. Enhancers were linked to their putative target genes using the GeneHancer72 database. To define actively transcribed regulatory elements, H3K27ac and ATAC-seq peaks under lupus-like condition were intersected (using bedtools112) with the promoters and enhancers as described above.

To uncover transcription factors that could potentially act in concert with SMC1A to regulate the genome during lupus inflammation, we performed motif enrichment analysis78 in chromatin regions with increased accessibility and SMC1A binding in lupus-like compared to basal monocytes (Supplementary Fig. 7). Among the top hits were STAT3, NF-κB, IRF3, and members of the ISGF3 (IFN-stimulated gene factor 3) complex –STAT1 and IRF9–79 suggesting their potential role in SMC1A-mediated reprogramming of monocytes in lupus.

SMC1A binding at active regulatory elements induces the transcriptional activation of inflammatory genes in lupus-like monocytes

Having observed transcriptional changes related to SMC1A dosage and a differential binding pattern of SMC1A during lupus inflammation, we asked whether changes in gene expression correlate with the activity of cognate enhancers and promoters and the chromatin enrichment or depletion of SMC1A. We used the log2FC values of lupus-responsive genes (Supplementary Data 1) to assess the level of transcriptional activation or repression according to the presence or absence of epigenomic features, based on our integrative analysis (Fig. 5A, B). In the absence of differential SMC1A chromatin binding, neither the activity of enhancers (Fig. 6A) nor of promoters (Fig. 6B) had a significant effect on the magnitude of gene activation during lupus condition (Fig. 6A, B: brown- versus white-colored bars). By contrast, increased binding of SMC1A on active regulatory elements was linked to significantly higher induction of the associated genes (Fig. 6A, B: red- versus brown-colored bars, p = 1.9e-5 for enhancers; p = 8.3e-06 for promoters). Transcriptional induction was even stronger for genes (n = 24) whose both active promoter and active enhancer exhibited increased binding by SMC1A (Fig. 6C). A similar analysis of genes repressed during lupus revealed no significant variation in log2FC values according to SMC1A binding (Supplementary Fig. 8A, B). These results suggest that transcriptional activation of genes correlates with enhanced binding of SMC1A to active chromatin, as opposed to gene repression, which is independent of SMC1A enrichment or depletion.

Fig. 6: SMC1A binding at active regulatory elements drives transcription of inflammatory genes in monocytes activated with lupus-relevant stimuli.
figure 6

A Bar-dot plot of expression changes (log2FC) in transcriptionally upregulated, lupus-responsive genes based on chromatin activity and SMC1A binding enrichment at their cognate enhancer regions in SLE-like versus unstimulated monocytes. Each dot represents a different gene. +/– signs indicate active/inactive chromatin, respectively and arrows indicate increased/reduced binding of SMC1A. Wilcoxon signed-ranked test was performed with multiple testing correction. Increased binding of SMC1A on active enhancers is associated with significantly higher induction of the associated genes (red- versus brown-colored bars, p = 1.9e-5). B Bar-dot plot of log2FC in transcriptionally upregulated, lupus-responsive genes based on chromatin activity and SMC1A binding enrichment at their cognate promoter regions in SLE-like versus unstimulated monocytes. Each dot represents a different gene. Wilcoxon signed-ranked test was performed with multiple testing correction. Increased binding of SMC1A on active promoters is associated with significantly higher induction of the associated genes (red- versus brown-colored bars, p = 8.3e-06). C Bar-dot plot of gene expression changes (log2FC) in transcriptionally upregulated, lupus-responsive genes based on enriched SMC1A binding at cognate active promoters (left bar), active enhancers (middle bar), or both (right bar). Each dot represents a different gene. In (AC), the box represents the interquartile range of log2FC values, with the vertical line indicating the median across each category of epigenome features. Whiskers extend to 1.5× the interquartile range, and points outside the whiskers represent outliers. Wilcoxon signed-ranked test was performed with multiple testing correction. D Effect of SMC1A gene silencing on log2FC of lupus-responsive genes, according to presence of active chromatin and enriched SMC1A binding at the cognate regulatory elements in SLE-like monocytes. +/– signs indicate active/inactive chromatin, respectively; E enhancer, P promoter. Filled circles represent means and error bars the 95% confidence interval. In each group of genes on the x-axis, one-samples t-test was performed to define statistically significant deviation from zero (** p < 0.01; *** p < 0.001). E Functional enrichment analysis (Gene Ontology) of Group B (n = 301) and C (n = 24) genes (D). F Functional enrichment analysis (Gene Ontology) of Group A genes (n = 530; Fig. 6D). All p-values are two-tailed.

To corroborate these results, we also examined the effect of SMC1A downregulation on the expression levels (log2FC values) of lupus-responsive genes according to the presence or absence of the aforementioned epigenomic features in their respective regulatory regions. Genes with combined enhancer/promoter activity and enriched SMC1A binding at the cognate enhancer only (Fig. 6D: group B, n = 301) showed an overall significant trend for downregulation upon SMC1A silencing. A more pronounced decrease was observed in genes with increased SMC1A binding at both their cognate active enhancers and promoters (Fig. 6D: group C, n = 24). Notably, these two groups of genes were enriched in inflammatory pathways, including cytokine signaling and response to virus (Fig. 6E). By contrast, genes with no active regulatory elements and no enriched binding by SMC1A (Fig. 6D: group A, n = 530), which also demonstrated reduced expression upon SMC1A silencing, participated in non-immune/inflammatory pathways, suggesting these might represent indirect effects (Fig. 6F).

Focusing on group B + C genes, we identified 277 genes that exhibited active enhancer, enriched SMC1A binding at the cognate enhancer and transcriptional upregulation (log2FC ≥ 1, adj. p < 0.05) under lupus conditions (Fig. , Supplementary Data 2). Enrichment analysis showed overrepresentation of cytokine-mediated signaling, response to LPS, response to virus, T-cell activation and positive regulation of cytokine production pathways (Fig. ). These genes may be considered putative direct targets of SMC1A as exemplified in the case of IL6 and GBP5 genes (Fig. 7C). Collectively, during lupus, SMC1A, potentially in collaboration with specific immune transcription factors, may be recruited at the active regulatory regions of immune-related and inflammatory genes, orchestrating their transcriptional regulation.

Fig. 7: SMC1A-regulated genes are involved in key inflammatory and immune pathways in SLE-like monocytes.
figure 7

A Volcano plot of DEGs belonging to Groups B and C (Fig. 6D) in SLE-like versus basal-state CD14⁺ monocytes. A subset of 277 genes met the criteria for SMC1A target genes, defined as showing significant transcriptional upregulation (log2FC ≥ 1, adjusted p < 0.05), active enhancer activity, and enriched SMC1A binding at the cognate enhancer. B Functional enrichment analysis (Gene Ontology terms) of the 277 SMC1A target genes. C Genome browser tracks showing SMC1A and H3K27ac ChIP-seq peaks in basal and SLE-like monocytes, together with ATAC-seq peaks in SLE-like monocytes, at the IL6 (left) and GBP5 (right) loci.

SMC1A-regulated inflammatory genes demonstrate female-biased expression in monocytes from SLE patients

In view of the above and considering that SMC1A exhibits female-biased expression in SLE, we sought evidence of its potential influence on sex-related transcriptional differences in patient monocytes. Blood was collected from a new cohort of female (n = 8) and male (n = 10) SLE patients, matched for age and disease characteristics (Supplementary Table 6), and gene expression profiling was performed in purified CD14+ monocytes by QuantSeq 3′ mRNA sequencing (GEO accession number GSE279728). We found a total of 745 genes with sexual differential expression (|log2-FC| > 1, p < 0.05) (Fig. 8A). Female-overexpressed genes in SLE monocytes (n = 596; Supplementary Data 3) showed overrepresentation of inflammatory pathways, including cytokine signaling, T-cell and immune response activation (Fig. 8B). On the other hand, genes with male-biased expression (n = 149; Supplementary Data 3) were enriched in ribosome biogenesis, cellular metabolism and mitochondrial function (Fig. 8C). Consistent with the gene expression data, female SLE monocytes secreted higher levels of the immune-response cytokines IL-6, IL-1b, TNF, and IL-10 as compared to their male counterparts (Fig. 8D). Together, our data underscore significant sexual differences in the molecular and inflammatory profile of SLE monocytes.

Fig. 8: Female-biased expression of SMC1A-regulated Inflammatory genes and cytokine release in SLE monocytes.
figure 8

A Volcano plot of DEGs from QuantSeq 3′ mRNA sequencing performed in blood monocytes from female (n = 8) compared to male (n = 10) SLE patients (|log2-FC| > 1, p < 0.05 generated using DESeq2). Blue dots represent genes overexpressed in males (n = 149), and red dots represent those overexpressed in females (n = 596) with SLE. B Functional enrichment analysis (Gene Ontology) of DEGs with female overexpression in SLE monocytes. C Functional enrichment analysis (Gene Ontology) of DEGs with male overexpression in SLE monocytes. D Detection of cytokines by LEGENDplex immunoassays at supernatants of cultured (4 h) SLE patient-derived monocytes (n = 9 pairs of males and females). Each dot represents an individual, the bar plots show the mean ± standard error (SE). Mixed model analysis was performed accounting for nested data structure and inter-individual variability, and introducing sex (male versus female) as within-cluster predictor. E Heatmap displaying z-score-transformed expression levels of 277 SMC1A-regulated genes in male and female SLE monocytes (same individuals as in (A); n = 8 females and n = 10 males). Each row represents a gene, and each column represents an individual. A total of 62 SMC1A target-genes (22.4%) were overexpressed in females, while only 2 (0.7%) were overexpressed in males with SLE (Fig. 8A-B). (F) Scatter plot showing the aggregated expression levels (after z-transformation) of all 277 SMC1A-regulated genes in female and male SLE monocytes (same individuals as in (E); n = 8 females and n = 10 males). Each dot represents a different individual and the bars indicate the minimum and maximum values. Mixed model analysis was performed as in (D). G Scatter plot showing the correlation between gene expression changes (Log2FC) in female versus male SLE monocytes (y-axis) and gene expression changes (Log2FC) in SMC1A si versus control si lupus-like monocytes (x-axis). Each circle represents a different protein-coding gene: red-colored are SMC1A target-genes whereas blue-colored are genes with active promoters and enhancers but without enriched SMC1A binding in SLE-like monocytes. Linear regression revealed negative associations which were stronger in SMC1A target-genes (β-coefficient ± SE = −1.861 ± 0.338, p = 8.1e-08) compared to non-SMC1A-target-genes (β ± SE = −0.673 ± 0.120, p = 3.1e-08) (p < 0.001 for the difference between the two slopes). All p-values are two-tailed.

Next, we focused on the 277 genes regulated by SMC1A in lupus environment (Fig. 4G, Supplementary Data 2) and compared their expression levels in female versus male SLE monocytes (Fig. 8E). First, we aggregated the expression levels (following z-transformation) of all SMC1A target genes in each SLE individual. Increased SMC1A regulon activity was noted in female than male monocytes (p = 0.0239) (Fig. 8F). In accordance, out of the 277 SMC1A target genes, 62 (22.4%) were overexpressed in females, while only 2 (0.7%) were overexpressed in males with SLE. Likewise, female-biased genes in SLE monocytes showed significant enrichment for those regulated by SMC1A (odds ratio 11.5, p = 2.1e-06, Fisher’s exact test). Notably, genes with stronger female-biased expression in SLE patients were more sensitive to SMC1A dosage in SLE-like monocytes (Fig. 8G).

Then, we explored whether these associations are specific to SLE by interrogating publicly available RNA-sequencing profiles from a large collection of healthy female and male monocytes20. Applying the same thresholds for sexually differential expression (|log2FC| > 1, adjusted p < 0.05) in both our SLE dataset and the aforementioned healthy dataset, we found that 51 out of 277 SMC1A-regulated genes were female-biased in SLE monocytes, but none in healthy monocytes. The opposite trend (SMC1A target genes with female-biased expression exclusively in the healthy state) was observed for only one gene. Collectively, these findings are consistent with SMC1A assuming a distinct regulatory role in lupus and suggest that its overexpression in female SLE monocytes may be linked to increased expression of its target genes, contributing to a more prominent inflammatory phenotype compared to male counterparts.

Female lupus-like monocytes exhibit elevated expression levels of IL6 and IL1A, along with increased SMC1A binding at their enhancers

We further investigated whether sex differences in SMC1A levels correlate with differential expression of its target genes by monitoring matched female and male CD14+ monocytes (n = 6 pairs) under unstimulated and stimulated conditions (sequential IFN-α/TNF and LPS, Fig. 1G). We selected two hallmark SMC1A-regulated genes, IL6 and IL1A, which are implicated in SLE autoimmunity.80,81 In patients with SLE, higher disease activity scores correlate positively with circulating IL-1α and IL-6 levels82,83, and serum IL-6 levels increase prior to SLE onset84. By RT-qPCR, both genes were significantly overexpressed in female compared to male monocytes under basal state (Fig. 9A, B). Following stimulation, both IL6 and IL1A were induced in either sex; however, female lupus-like monocytes had significantly higher transcript levels compared to male counterparts.

Fig. 9: Female lupus-like monocytes show increased expression levels of IL6 and IL1A due to enhanced SMC1A occupancy at their cognate enhancers.
figure 9

RT-qPCR was performed to quantify IL6 (A) and IL1A (B) mRNA in CD14+ monocytes (n = 6 pairs of male and female donors) cultured under unstimulated (basal state) and SLE-like conditions. Quantification was performed using the 2–ΔΔCT method. Each dot represents a different donor, and bar plots show the mean ± standard error (SE). Mixed model analysis was performed accounting for nested data structure and inter-individual variability, and introducing sex (male versus female), monocyte condition (SLE-like versus basal) and the interaction of sex × monocyte condition as within-cluster predictors. The Least Squares Means (LSM) method was used for post-hoc comparisons between selected groups. RT-qPCR was performed to quantify IL6 (C), IL1A (D) and SMC1A mRNA levels in monocytes (n = 4 pairs of male and female donors) cultured under unstimulated (basal state) and SLE-like conditions. Mixed model analysis was performed with IL6 (C) or IL1A (D) as the dependent variable, using sex, monocyte condition, SMC1A expression levels and their interaction with monocyte condition as within-cluster predictors. The corresponding beta (β) coefficients and p-values are shown. E ChIP-PCR for chromatin-normalized SMC1A enrichment (relative to gene desert region) at enhancer regions of IL6 (left panel) and IL1A (middle panel) and promoter region of IL6 (right panel) in lupus-like monocytes (n = 4 pairs of male and female donors). Each dot represents a different donor, and bar plots show the mean ± SE enrichment. Mixed model analysis was performed accounting for nested data structure and inter-individual variability, and introducing sex as within-cluster predictor. F RT-qPCR was performed to measure IL6 and IL1A mRNA in SMC1A-silenced (SMC1A si) compared to control-silenced (ctrl si; scramble) monocytes cultured under lupus conditions (n = 4 replicates). Quantification was performed using the 2–ΔΔCT method. Each dot represents a different donor, and bar plots show the mean ± SE. Paired t-test was performed. G ELISA was performed to quantify secretion of IL-6 by SMC1A-silenced (SMC1A si) compared to control-silenced (ctrl si; scramble) monocytes cultured (4 h) under lupus conditions (n = 9 replicates). Each dot represents a different donor, and bar plots show the mean ± SE. Paired t-test was performed. All p-values are two-tailed.

We then employed mixed model analysis to investigate whether the transcriptional levels of these genes could be predicted by SMC1A abundance, and whether this relationship is influenced by donor sex and/or the underlying cell state (lupus-like versus basal). SMC1A mRNA significantly predicted the expression of both IL6 (p < 0.001) and IL1A (p < 0.001) (Fig. 9C, D). This effect was more pronounced than that of sex (i.e., higher beta coefficients of SMC1A). These data support the positive regulatory role of SMC1A on IL6 and IL1A and suggest that sex differences in the expression of these genes may be explained –at least to some extent– by variations in SMC1A levels.

To obtain more direct evidence linking sexual differences in SMC1A with the transcription of its target genes, we quantified SMC1A binding at the regulatory elements of IL6 and IL1A in female versus male lupus-like monocytes. Utilizing our ChIP-seq data from male CD14+ monocytes (Fig. 4), we defined the active chromatin regions –marked by H3K27Ac and ATAC-seq peaks– with increased SMC1A binding during lupus inflammation at the enhancer and promoter of IL6 and the enhancer of IL1A. Following ChIP experiments in matched female and male donors, we observed SMC1A enrichment at the IL6 and IL1A enhancers, thus confirming our ChIP-seq data. In both elements, fold enrichment of SMC1A (over a negative control genome region) was significantly higher in female compared to male lupus-like monocytes (Fig. 9E). These data suggest that female overexpression of IL6 and IL1A in the lupus environment might be facilitated by enhanced SMC1A binding at the cognate active enhancers, highlighting putative biological consequences of sex variation in SMC1A abundance. As a functional confirmation, lowering SMC1A in SLE-like monocytes decreased the mRNA levels of IL6 and IL1A (Fig. 9F) as well as the secreted levels of IL-6 (Fig. 9G).

Discussion

Although certain mechanisms explaining the predominance of autoimmunity in women have been proposed7,27, the molecular basis for this bias remains incompletely understood. Here, we implicate the cohesin complex member SMC1A, a chromatin architectural factor that escapes XCI, in sex-divergent inflammatory responses in the context of lupus. Using an unbiased approach, we identified monocytes as a pathophysiologically relevant immune cell type exhibiting exaggerated enhanced female-biased SMC1A expression in both patients and an ex vivo model of SLE, compared to healthy individuals and those with AS. Under lupus conditions, genes with deregulated expression are particularly sensitive to SMC1A dosage and mechanistically, SMC1A assumes a specific regulatory role by repositioning to the active, accessible enhancers of a subset of immune and inflammatory genes, driving the transcriptional reprogramming of monocytes. IL6 and IL1A, two SMC1A regulated-genes with an important role in autoimmunity, are significantly elevated in female SLE-like monocytes as opposed to male ones. This increase in expression may be attributed to sexual differences in SMC1A abundance and its binding at the respective enhancers. Importantly, the female-biased transcriptome in SLE patient-derived monocytes shows significant enrichment for SMC1A-regulated genes involved in inflammatory processes, therefore highlighting the role of this sex-biased genome organizer in predisposing to autoimmunity.

A key observation in our study was the prominent female-bias of SMC1A expression in blood immune cells, particularly monocytes, of SLE compared to healthy individuals. While we did not examine for SMC1A biallelic patterns in these samples, our finding is consistent with the concept proposed by Anguera and colleagues that SLE female immune cells display abnormal maintenance of XCI due to paucity of repressive epigenetic marks on the inactive X chromosome7,85,86,87. This enables the re-activation of escapee genes leading to increased mRNA levels as illustrated for CD40LG87. Of note, SMC1A tended to be downregulated in SLE and after exposure to lupus milieu, possibly due to a negative feedback mechanism. Even under these conditions, females retained significantly higher expression of SMC1A than males, a pattern not observed in AS. Whether this discrepancy between the two disorders is due to differences in genetic20, cytokine (type I IFN in SLE61, IL-23/IL-17 in AS88) or epigenetic landscapes, remains to be defined. These findings support the relevance of our approach in systematically exploring sexual genomic divergences not only in the normal state but also in the context of sex-biased immune disorders.

Sex-related molecular differences manifest cell-type specificity15,19,20 and may be particularly prominent in monocytes/macrophages19,20,22,23,60. This is reiterated by our studies in SLE-like monocytes, where SMC1A exhibits sex-dependent chromatin interactions linked to gene expression. Monocytes are pivotal to SLE disease with multifaceted roles in sensing and reacting to danger signals, promoting autoreactive lymphocytes and contributing to organ injury32,38,89. Our evidence further indicates that these cells can shape sexual inflammatory responses in lupus. In fact, through a comparative transcriptome analysis in patient-derived monocytes, we herein define the molecular fingerprints of female and male SLE, thus enhancing our understanding of the susceptibility and heterogeneity of this complex disease. Moreover, these data could provide a platform for developing personalized treatment strategies to overcome sex disparities in the effectiveness of immunotherapies90.

Consistent with the regulatory role of the cohesin complex46, we demonstrated that SMC1A accumulates at the active genome elements of lupus-responsive genes in monocytes. This agrees with a study in retinal pigment epithelium where cohesin (Rad21) bound at highly active chromatin enriched for H3K27ac and H3K4me3 marks91. Besides acting as an insulator protein (when colocalized with CTCF) at topologically associating domain (TAD) boundaries, cohesin may directly control transcription by multiple mechanisms such as facilitating enhancer–promoter connections, collaborating with RNA polymerase II92 and/or tissue-specific transcription factors50,55. Despite the overall reduced SMC1A binding at promoter regions during transition from naïve to lupus-like state, we observed the strongest induction in genes with concerted SMC1A enrichment to both active enhancers and cognate promoters. These genes were particularly influenced by SMC1A downregulation pointing towards direct -rather than secondary- effects. Although we did not map the genome structure, our findings align with those of Robles-Rebollo et al.93, who proposed that cohesin links the probabilities of enhancer and target gene bursting in stimulated murine macrophages, an observation not entirely explained by spatial proximity. Additionally, our data reiterate the importance of cohesin presence at transcription start sites for gene expression91,94.

Cohesins are linked to cell- and context-specific genome effects46,91 and we herein establish a critical role for SMC1A in the inflammatory gene program of human monocytes under steady-state and lupus conditions. This agrees with observations that acute myeloid leukemia cells bearing cohesin mutations exhibit impaired abundance of immune/inflammation-related genes56,57. Likewise, macrophages and hematopoietic stem cells (HSC) from Rad21-null or heterozygous mice exhibit a blunted inducible inflammatory response, attributed to disturbed accessibility and activation of cognate enhancers enriched with IFN-stimulated response elements and IRF-PU.1 composite motifs57 and uncoupled burst frequencies of inducible enhancers and promoters93. Studies in murine HSCs have also highlighted that Rad21 mediates NF-κB signaling by inducing the enhancer activity of co-transcription factors such as ETS, AP1 and Runt family members56. To this end, our in-silico analyses suggest that the specificity of SMC1A-mediated inflammatory reprogramming in SLE-like monocytes is possibly facilitated by the engagement of key immune-related transcription factors, including STAT3, NF-κB, STAT1, IRF3 and IRF9. These findings are consistent with work in LPS-stimulated macrophages revealing that cohesin-dependent inducible enhancers show enriched binding of STAT2 and IRF393. Moreover, STAT1 and IRF9 form a complex with STAT2 to create the ISGF3 complex79 which regulates the type I IFN transcriptional response. Notably, the aforementioned factors play significant roles in the immune function including their involvement in autoimmune disorders like SLE95,96.

Previous research has provided a mechanistic basis for how certain X-linked genes with incomplete dosage compensation may account for female susceptibility to SLE. One example is TLR7, which is overexpressed in SLE and whose biallelic expression correlates with augmented type I IFN synthesis by plasmacytoid dendritic cells97,98 and the maturation of memory B-cells11. A second is CXorf2110 encoding for the SLC15A4-associated adapter that mediates TLR7/9-induced IRF5 activation99. Our work implicates SMC1A as a master genome regulator that shapes the expression of multiple immune and inflammation-related genes in monocytes. Female-biased expression of SMC1A was linked to heightened inflammatory responses, exemplified by the SMC1A-regulated genes IL6, IL1A, and GBP5, which showed female overexpression in SLE patients or in monocytes cultured under lupus-like conditions. Of note, IL6 and GBP5 have been previously classified as female-biased genes in healthy monocytes20 and in the whole blood upon influenza vaccination100. We demonstrate that the higher expression of IL6 and IL1A in female than male SLE-like monocytes may be driven –at least to some extent– by augmented SMC1A binding at the corresponding enhancers in the former group.

To this end, we propose a mechanism by which failure of female SLE monocytes to sufficiently downregulate SMC1A in response to SLE-like inflammation may contribute to exaggerated inflammatory gene expression (Fig. 10). This concept is further supported by observations that aged hematopoietic stem and progenitor cells exhibit impaired downregulation of Rad21/cohesin, which hinders their ability to constrain NF-κB-mediated inflammation56. Nonetheless, we cannot rule out the possibility that the female propensity to SLE is driven also by sex differences in SMC1A expression in other immune and non-immune26 cell types.

Fig. 10: Proposed mechanism of female-biased SMC1A expression in shaping the inflammatory molecular program of monocytes and predisposition to lupus autoimmunity.
figure 10

SMC1A, a cohesin complex subunit escaping X-chromosome inactivation, shows enhanced female-biased expression in monocytes from SLE patients as well as in those cultured under lupus-inducing conditions. Under lupus-inducing conditions, SMC1A is redistributed to the active regulatory elements of immune-related genes in monocytes, leading to increased transcriptional levels. Increased SMC1A binding at the enhancers of inflammatory genes, such as IL6 and IL1A, in female than male monocytes during lupus may lead to higher expression levels in the former group. Importantly, SMC1A-regulated genes are upregulated in female monocytes from SLE patients, correlating with a more inflammatory molecular and cytokine fingerprint, as compared to male counterparts. Created in BioRender.

In conclusion, our work highlights the role of SMC1A/cohesin as an X-linked chromatin modifier that shapes a genome environment permissive to inflammatory gene expression, shedding light on the molecular basis of sexual differences in immune responses and providing a rationale for relevant chromosome conformation studies. Increased abundance and interaction of SMC1A with active enhancers culminate in augmented levels of inflammatory genes and soluble mediators by female monocytes exposed to lupus-inducing stimuli, therefore rendering them prone to autoimmunity. SMC1A and its target genes represent potential biomarkers for risk stratification and/or druggable targets for the prevention and treatment of SLE and related female-biased immune pathologies.

Methods

Patients, healthy donors, and ethics committee approval

Patients with SLE were recruited from the Rheumatology Clinics of the University Hospital of Heraklion and the “Attikon” University Hospital, meeting the 2019 EULAR/ACR classification criteria101 (Supplementary Tables 1, 2, 4, and 6). SLE disease activity was assessed using the validated SLEDAI-2000 (SLEDAI- 2K)102, with low activity defined according to the Lupus Low Disease Activity State (LLDAS)103. Patients with axial spondyloarthritis, classified per ASAS criteria104, were also recruited from the University Hospital of Heraklion, and their disease activity was measured using the ASDAS-CRP index105. All patients abstained from medications for at least 12 h prior to blood collection. Healthy blood samples were obtained from the Blood Transfusion Unit of the University Hospital of Heraklion. The study was approved by the Ethics Committees of the University Hospital of Heraklion (protocol no. 5944/14-6-2017) and the Attikon University Hospital (protocol no. 299/3-6-2020). All participants provided written informed consent.

Peripheral blood immune cell subtypes isolation

Peripheral blood mononuclear cells were isolated from heparinized blood by density-gradient centrifugation using Lymphocyte Separation Media (Biowest, #L0560). Erythrocytes were removed by hypotonic lysis (1 ml ddH2O for 25 s followed by 1 ml 1.8% NaCl). Cell viability was consistently >95%, as measured by trypan blue dye (Gibco, #15250-061) exclusion in all experiments. CD14+ monocytes and CD4+ T-lymphocytes were magnetically isolated by positive selection, while CD19 + B-lymphocytes were isolated by negative selection (MACS, Miltenyi Biotec, #130-050-201, #130-045-101, #130-091-151). The purity of all cell types was confirmed by flow cytometry (CD14+, CD4+, CD19+ cells/total live cells), and only preparations with >94% purity were used. Peripheral blood neutrophils were isolated from heparinized human blood using a histopaque double-gradient density centrifugation (Sigma, #11191).

Monocyte culture

Magnetically isolated CD14+ monocytes were cultured in RPMI-1640, GlutaMAX Supplement (Gibco, #61870-010), including 10% heat-inactivated FBS (Gibco, #10437028), 100 IU/mL penicillin and 100 μg/mL streptomycin (Gibco, #15070063). To simulate SLE, CD14+ monocytes isolated from healthy donors were stimulated with 10 ng/ml human macrophage colony-stimulating factor (M-CSF; Peprotech, #300-25), 10 ng/ml recombinant human TNF (Peprotech, #300-01Α), 100 IU/ml IFN-α (PBL Science, #PBL11200-2) overnight at 37 °C/5% CO2, followed by 10 ng/ml LPS (Sigma, #L4391) challenge for 1–4 h68. In other experiments, monocytes obtained from healthy donors were cultured in presence of serum derived from SLE patients or healthy donors (20% v/v) with or without 17-beta-estradiol (10−9M; kindly provided by Prof. E. Castanas, University of Crete) for 18 h at 37 °C/5% CO2.

Flow cytometry

Cells were stained for membrane markers for 20 min at 4 °C in PBS (Biosera, #LM-S2041) and 5% FBS. Conjugated antibodies against CD14, HLA-DR and CD80 were from BioLegend (#325604, #307628 and #305220, respectively). For apoptosis detection, cultured monocytes were stained with annexin V (BioLegend, #640922) and 7-AAD (Biolegend, #420404) according to the manufacturer’s recommendations. Samples were acquired on a BDFACSCanto II (BD Biosciences) and analyzed using the FlowJo software (Tree Star).

Enzyme-linked immunosorbent assay

Human IL-6 (R&D, #DY206-05), IL-1B (Invitrogen, #88-7261-88), and CXCL10 (Invitrogen, #BMS284INST) levels in culture supernatants, harvested at the indicated time points, were measured using sandwich ELISA following the manufacturer’s instructions. Light absorbance was recorded with an ELx800 microplate reader (Biotek). All samples were analyzed in duplicates (technical replicates).

TaqMan gene expression assay

Total RNA was isolated from cells using the NucleoSpin RNA extraction kit (Macherey-Nagel, #740955) following the manufacturer’s instructions. First-strand complementary DNA (cDNA) was synthesized using the PrimeScript RT-PCR Kit (Takara, #RR037A). DNA samples (50 ng/reaction) were combined with TaqMan Universal Master Mix II, no UNG (Applied Biosystems, #4440040), and reactions were performed using the CFX Connect Real-Time System (BioRad). Each multiplex RT-PCR reaction had a total volume of 20 μl. The SMC1A primers and probe (Hs00196849_m1) and GAPDH primers and probe (Hs02786624_g1) were part of TaqMan gene expression assays from Applied Biosystems.

Quantitative polymerase chain reaction

Total RNA from monocytes was collected using the TRIzol reagent (Invitrogen, #15596026) and TURBO DNA-free Kit (Invitrogen, #AM1907,) according to the manufacturer’s instructions. First-strand complementary DNA synthesis was performed using the PrimeScript RT-PCR Kit (Takara, #RR037A). qPCR was carried out using the KAPA SYBR Fast Universal Kit (Kapa Biosystems, #KK4602) at a CFX Connect, Real-Time System (Bio-Rad) or a Corbett Research Rotor-Gene 6000 Real-Time PCR Detection System (Qiagen). Expression was normalized to 18S and GAPDH and calculated by the change-in-threshold method (2–ΔΔCT). The primer sequences used were the following: SMC1A forward, 5′- CATCAAAGCTCGTAACTTCCTCG −3′; SMC1A reverse, 5′- CCCCAGAACGACTAATCTCTTCA −3′; 18S forward, 5′- CTCAGCGTGTGCCTACCCTA −3′; 18S reverse, 5′- GCAAGCTTATGACCCGCACT − 3′; GAPDH forward, 5′- CATGTTCCAATATGATTCCACC −3′; GAPDH reverse, 5′- GATGGGATTTCCATTGATGAC −3′; IL6 forward, 5′- CAGATGAGTACAAAAGTCCTGA −3′; IL6 reverse, 5′- CTACATTTGCCGAAGAGCCC −3′; IL1A forward, 5′- CGCCAATGACTCAGAGGAAGA −3′, IL1A reverse, 5′- AGGGCGTCATTCAGGATGAA −3′; GBP5 forward, 5′- CACGAAACTGGTCCCTGATG −3′; GBP5 reverse, 5′- TACCCCTTAGCCTGTAAGCC −3′; IFNB forward, 5′- AGCTCCAAGAAAGGACGAACA −3′, IFNB reverse, 5′-GCCCTGTAGGTGAGGTTGAT −3′; CXCL10 forward, 5′- GTGGCATTCAAGGAGTACCTC −3′; CXCL10 reverse, 5′- TGATGGCCTTCGATTCTGGATT −3′; SLPI forward, 5′- CCCAAACCCAACAAGGAGGA −3′; SLPI reverse, 5′- GCAACACTTCAAGTCACGCT −3′.

Immunoblotting

Whole-cell extracts were lysed using RIPA Lysis Buffer (Millipore, #20-188,) supplemented with protease and phosphatase inhibitors (Roche, #04693124001, #04906837001) by vortexing on ice (every 5 min for 20 min with 1-min breaks). The lysates were centrifuged at maximum speed for 10 min at 4 °C, and the supernatant (protein lysate) was collected. Protein concentrations were measured using the Pierce BCA protein assay kit (Thermo Scientific, #23227). β-Mercaptoethanol (6×) was added to the samples followed by heating for 5 min at 95 °C. Equal amounts of protein (20 μg per well) were loaded onto 10% SDS-PAGE gels and transferred to Immobilon-PSQ PVDF membranes (Millipore, #ISEQ00010,) using a transfer apparatus (BioRad). Membranes were blocked with 5% BSA (BioFroxx, #9048-46-8) in tris-buffered saline with 0.1% Tween 20 (TBS-T) for 1 h, followed by overnight incubation at 4 °C with primary antibodies and a 1-h incubation with HRP-conjugated anti-mouse or anti-rabbit secondary antibodies. Blots were developed using enhanced chemiluminescence (Millipore, #WBLUF0100) and detected with an ImageBlot system (BioRad). Band intensities were quantified using Image Lab software, with total protein levels normalized to housekeeping protein expression. The primary antibodies used were: anti-SMC1A (Bethyl Laboratories, #A300-055A), anti-GAPDH (Invitrogen, #39-8600), and anti-ACTIN (Millipore, #MAB1501). Secondary antibodies used were anti-mouse HRP (Millipore #12-349) and anti-rabbit HRP (Invitrogen, #31460). Information on the dilutions used can be found in Supplementary Table 7.

Electroporation of primary monocytes with siRNAs

Blood CD14+ monocytes isolated from healthy donors were incubated overnight in RPMI-1640, GlutaMAX Supplement including 10% heat-inactivated FBS, 100 IU/mL penicillin, 100 μg/mL streptomycin and 10 ng/ml MCSF, in a humidified 37 °C/5% CO2 incubator. 106 cells/reaction were electroporated with 100 nM ON-TARGETplus SMC1A siRNA (Dharmacon, #L-006833-00-0005; SMC1A si monocytes) and ON-TARGETplus Non-targeting siRNA (Dharmacon, #D-001810-01-05; ctrl si monocytes) using Amaxa Human Monocyte Nucleofector Kit (Lonza Bioscience, #VPA-1007) and program Y-010. Cells were incubated with transfection mix in growth medium for 4 h and then were washed and incubated for 1–2 days with IFNα, TNF and LPS (see above). Real-time RT-qPCR and Western immunoblotting were used to evaluate silencing efficiency.

Chromatin immunoprecipitation

Blood CD14+ monocytes (30–50 × 10⁶), isolated from healthy donors, were crosslinked at room temperature for 10 min by adding 1% formaldehyde (ThermoScientific, #28906) to the growth medium. The crosslinking solution also contained 50 mM HEPES (pH 7.5), 100 mM NaCl, 1 mM EDTA (pH 8.0), and 0.5 mM EGTA (pH 8.0). After 10 min, crosslinking was quenched by the addition of 125 mM glycine for 10 min. Cells were pelleted at 4 °C and washed with ice-cold PBS. The crosslinked cells were lysed on ice for 10 min in lysis buffer (50 mM HEPES-KOH pH 7.5, 140 mM NaCl, 1 mM EDTA, 10% glycerol, 0.5% NP-40, and 0.25% Triton X-100) supplemented with 1X protease/phosphatase inhibitors. The lysates were centrifuged at 2000 × g for 5 min at 4 °C. The pellet was resuspended and sonicated in shearing buffer (50 mM Tris-HCl pH 8.0, 10 mM EDTA, 0.1% sodium deoxycholate, and 1X protease/phosphatase inhibitors) using a Bioruptor (Diagenode) with 30 s ON and 30 s OFF at high power output for 20 cycles. After sonication, samples were centrifuged at 15,000 × g for 10 min at 4 °C, and a small aliquot of the sonicated extract was saved as input. The remaining extract was incubated with Dynabeads Protein A and G (Invitrogen, #10001D/#10003D) for 30 min at 4 °C. Precleared extracts were then incubated with 90 μL Dynabeads (45 μL each of Dynabeads A and G) along with 2 μg or 10 μg of the appropriate antibody (specified below) overnight at 4 °C. ChIP-grade antibodies used were anti-H3K27ac (Abcam, #ab4729) and anti-SMC1A (Bethyl Laboratories, #A300-055A). Following overnight incubation, antibody-bound Dynabeads were washed once with low-salt wash buffer (0.1% SDS, 1% Triton X-100, 2 mM EDTA, 20 mM Tris-HCl pH 8.1, 150 mM NaCl), once with high-salt wash buffer (containing 500 mM NaCl), once with LiCl wash buffer (10 mM Tris-HCl pH 8.0, 1 mM EDTA, 250 mM LiCl, and 1% NP-40), and twice with TE buffer. Cross-links were reversed by overnight incubation at 65 °C. RNA and proteins were digested using RNase A and Proteinase K (Thermo Scientific, #EN0531, #EO0491), respectively, and DNA was purified using the MinElute PCR Purification Kit (Qiagen, #28004).

Chromatin immunoprecipitation coupled with polymerase chain reaction

Immunoprecipitated DNA was analyzed by real-time RT-qPCR and results were calculated using the percent input method (Thermo Scientific) and normalized to a negative control region (gene desert). The primer sequences used for ChIP-PCR were purchased by Eurofins Genomics and include: IL6 TSS forward, 5′- TCCAGCCCAGCATTAACAAG −3′; IL6 TSS reverse, 5′- CCAAAGCCCAACAGATGCTA −3′; IL6 enhancer forward, 5′- GGTGGATTGACGGGATCTGA −3′; IL6 enhancer reverse, 5′- GGGCTGAGGTGGGATAATCA −3′; IL1A enhancer forward, 5′- CAGTTCCATGCAAACCTGCC −3′; IL1A enhancer reverse, 5′- TCTTTGCCACCTCATTCCCC −3′; gene desert forward, 5′- ATGTCAGGCCCATGAACGAT −3′; gene desert reverse, 5′- GCATTCATGGAGTCCAGGCTTT −3′

ChIP-sequencing library preparation

Immunoprecipitated and input DNA samples were quantified with the qubit fluorometric method (Thermo Scientific) and ChIP-seq NGS libraries were prepared at the Greek Genome Center of the Biomedical Research Foundation Academy of Athens (BRFAA)106. Quantitation and quality control of the libraries was performed with Qubit HS DNA kit (Invitrogen, #Q32854) and Agilent Bioanalyzer DNA1000 kit (#5067-1504) respectively. Approximately 25 M, 100 base pair (bp) long, single-end reads were produced per sample, in the Illumina NovaSeq 6000 sequencer.

ChIP-sequencing pre-processing

ChIP-seq raw fastQ files were pre-processed by first assessing quality control (QC) with FASTQC. Low-quality 3′ read ends (minimum two-color chemistry score of 20), N tails, and adapter remains were trimmed/clipped using cutadapt107, allowing reads of at least 25 nucleotide length. Filtered reads were then mapped against the UCSC hg19 human reference genome using bwa-mem aligner108 with default parameters. Mapped reads were further filtered using samtools109, in order to discard chimeric alignments, alignments with more than 2 mismatches, and alignments with MAPQ quality lower than 10. A deduplicated version of alignments was also created using samtools markdup command (version 1.10), in order to eliminate the effect of PCR over-amplification during visualization. Alignment QC was further accessed using deeptools110. Specifically, biological replicates and conditions were examined regarding their similarity/dissimilarity with Pearson’s R correlation analysis (pairwise Pearson’s R metrics) and Principal Component Analysis (PCA), using genomic bin counts of 1 kb and 3 kb windows. Visual inspection of ChIP-seq signal was also applied by generating UCSC genome browser tracks111 (generation of bigWib files with deeptools bamCoverage110).

Assay for transposase-accessible chromatin-sequencing pre-processing and analysis

ATAC-seq raw FASTQ files were downloaded from the NCBI GEO database (accession code GSE100383) using the SRA toolkit (version 2.9.6; https://hpc.nih.gov/apps/sratoolkit.html), specifically the prefetch command. The FASTQ files were concatenated by biosample ID and processed for quality control (QC), adapter clipping, quality trimming, mapping, and alignment filtering, as described above. Peak files for each biological replicate (n = 5 unstimulated monocytes; n = 3 lupus-like monocytes) were also downloaded from the GEO repository. Replicates of both alignments (BAM files) and peaks (BED files) were merged for visualization purposes using samtools109 and bedtools112 (version 2.30), respectively. Merged peaks were further processed with bedtools merge to eliminate overlapping peak regions.

ChIP-sequencing peak calling

For each SMC1A ChIP-seq IP-INPUT pair, peaks were called by macs2113 with the following custom parameters enabled: “--nomodel”, “--shift 100”, and “--extsize 200”. H3K27ac ChIP-seq peak calling was performed by epic2114, an ultra-performant reimplementation of the SICER peak caller115, keeping duplicates (-kd parameter), using a bin of 400 base pairs (bps), allowing 1 gap, and a fragment size of 150 bps. Peaks spanning ENCODE blacklisted regions116 or including less than 10 mapped reads were excluded.

Differential peak analysis

Differential binding analysis for SMC1A ChIP-seq and differential accessibility analysis for ATAC-seq were conducted with the diffBind R package (version 2.10)117 using the default peak merging strategy of the tool as a reference, disabling summit calling, enabling TMM normalization, and using DESeq2118 for statistical testing. Analysis of differential H3K27ac enrichment was performed with a later version of diffBind (version 3.10). Differentially enriched regions were detected (DESeq2), and filtered by applying log2-fold change (log2FC) ≥ 1 and p ≤ 0.001 thresholds respectively for SMC1A and H3K27ac ChIP-seq assays, and log2FC ≥ 3 and p ≤ 0.001 thresholds for ATAC-seq. Alternatively enriched regions were summarized with custom-generated volcano plots using R. Peaks were then annotated regarding their genomic position, using ChiPSeeker119. Specifically, the genome locations of the peaks were assigned to different transcript-related features (such as promoter, 5′ UTR, 3′ UTR, exon, intron, distal intergenic region) and the nearest gene, obtained from TxDb.Hsapiens.UCSC.hg19.knownGene. The euler diagram showing gene sets with differential binding of SMC1A, differential enrichment of H3K27ac and shared genes within each set was generated with the R package eulerr (version 7.0.2). For this analysis, only the nearest peak region to the transcription start site was retained for each gene, to avoid mixing up- and down-bound peaks for the same gene.

RNA sequencing

RNA was isolated from unstimulated (basal-state) and stimulated (SLE-like) monocytes (ctrl si or SMC1A si), using the NucleoSpin RNA extraction kit (Macherey-Nagel, #740955) according to manufacturer’s instructions. RNA was quantitated with the Nanodrop photometric method and quality control was performed with the Agilent Bioanalyzer RNA nano kit (#5067-1511). Libraries were prepared with Illumina TruSeq RNA library prep kit (#RS-122-2001) at the Greek Genome Center of BRFAA. Approximately 25 M, 100 base pair (bp) long, single-end reads were produced per sample, in the Illumina NovaSeq 6000 sequencer.

QuantSeq 3’ mRNA sequencing

Βlood-derived CD14+ monocytes of 10 male and 8 female SLE patients cultured for 18 h with 10 ng/ml M-CSF (Peprotech, #300-25). Supernatants were collected for measurement of cytokines and cell pellets for RNA isolation using the NucleoSpin RNA extraction kit (Macherey-Nagel, #740955). RNA samples underwent quality control, and NGS libraries were generated using the QuantSeq 3′ mRNA-Seq Library Prep Kit FWD (Lexogen, #191.96), according to the manufacturer’s protocol, at the Genomics Facility of IMBB (FORTH, Heraklion). High-throughput sequencing (75 bp, single-end) was performed on an Illumina platform (NextSeq 500), following the manufacturer’s recommended protocol.

RNA-sequencing and Quant-sequencing data analysis

Raw fastQ files were preprocessed with the same quality trimming and adapter clipping strategy as described above for ChIP-seq preprocessing. Filtered reads were then mapped against the UCSC hg19 human reference transcriptome using STAR aligner120 by setting the “--outFilterMultimapNmax” parameter to 5. As with ChIP-seq data, mapped reads were filtered in order to discard chimeric alignments, alignments with more than 2 mismatches, and keep only uniquely aligned reads (MAPQ = 255). Counting on Ensembl genes and statistical analysis (differential expression analysis) was performed with metaseqR121 using DESeq2118 for normalization and differential testing, and applying default settings for each assay (exon count summarization for RNA-seq and 3’ UTR summarization for Quant-seq). For paired-samples comparisons, “donor id” was included in the DESeq2 model.

Integrative analysis of ChIP-sequencing, ATAC-sequencing and RNA-sequencing data

Differentially SMC1A-bound regions were intersected with Ensembl Transcription Start Sites (TSSs; a total of 57,735 records from the July 2020 version) and FANTOM572 enhancer elements (enhancer TSSs [eTSSs]; 65,423 human enhancers from phases 1 and 2 of the project, derived from http://fantom.gsc.riken.jp/5/datafiles/phase2.2/extra/Enhancers/human_permissive_enhancers_phase_1_and_2.bed.gz), each extended by 500 base pairs, upstream and downstream. Only promoters from transcribed genes, including protein-coding genes, lincRNAs, miRNAs, and lncRNAs were kept in the dataset (pseudogenes and other dubious ORFs were excluded). Enhancers were linked to their putative target genes using the GeneHancer72 database, which provides enhancer-gene associations. To define actively transcribed regulatory elements, H3K27ac and ATAC-seq peaks under lupus-like condition were intersected (using bedtools112) with the promoters and enhancers as described above.

Visualization

SMC1A ChIP-seq H3K27ac ChIP-seq and ATAC-seq signals were summarized (bin-counts) and visualized across active promoter regions (2 kb around active TSSs) and active enhancer regions (2 kb region eTSSs) as heatmaps and average profiles, using seqMINER122 and custom R scripts. Replicates for each condition were first downsampled to the same alignment depth, and then merged accordingly. Cross-condition normalization was also performed with respect to the alignment depth.

Generation of UCSC genome browser tracks

To generate genome-wide views of NGS signal, for all assays described in this study, deeptools bamCoverage110 was utilized, with a bin size of 20, CPM normalization, exactScaling, and the default effectiveGenomeSize for hg19 reference. This procedure generated bigWig signal files that were used to create tracks comparable with the UCSC genome browser.

Gene Ontology and motif enrichment analysis

To identify enriched biological processes, molecular functions, and cellular components associated with our gene sets, we conducted Gene Ontology (GO) enrichment analysis using the enrichGO function from the Clusterprofiler123 package. The analysis was performed with the following parameters: genes were annotated using the “SYMBOL” key type and mapped to the human gene annotation database org.Hs.eg.db. The Benjamini-Hochberg method was applied to adjust p-values for multiple comparisons and enrichments. Motif enrichment analysis was performed using the i-cisTarget78, screening for chromatin regions with increased accessibility and SMC1A binding in lupus-like compared to normal monocytes.

Measurement of secreted cytokines

Secreted cytokines from the supernatants of SLE monocytes, cultured for 18 h with 10 ng/ml M-CSF (Peprotech, #300-25), were measured with the Human Anti-Virus Response Panel (13-plex) from BioLegend (#740349) according to the manufacturer’s instructions. Data acquisition was done on BDFACSCanto II (BD Biosciences) and analyzed using the LEGENDplex Data Analysis Software (https://legendplex.qognit.com/).

Statistical analysis

Data were analyzed using STATA (version 18.5, StataCorp, Texas, USA) or GraphPad Prism 9 software. Results are expressed as median or mean ± standard error (SE). Statistical analysis was conducted using paired or independent-samples t-tests, as appropriate. For assays involving clusters of female and male samples (patients, healthy donors, lupus-like monocytes), linear mixed models (GLMM) were employed to account for the nested data structure and inter-individual variability. In these models, sex (female versus male) was treated as a within-cluster predictor alongside other covariates. Depending on data skewness, either a Gamma or Gaussian distribution was applied, with a log- or identity-link function, respectively. Post-hoc comparisons were performed using Tukey’s Least Significant Difference (LSD) test. The relationship between SLE disease activity (SLEDAI-2K) and SMC1A expression levels in patient-derived CD14⁺ monocytes was assessed by linear regression. Two-tailed p-values are reported.