Abstract
   Osteoarthritis affects millions worldwide, yet effective treatments
   remain elusive due to poorly understood molecular mechanisms. While
   genome-wide association studies (GWAS) have identified hundreds of
   osteoarthritis-associated loci, identifying the genes impacted at each
   locus remains challenging. We investigate alternative splicing using
   RNA-sequencing data from 101 human chondrocyte samples treated with
   phosphate-buffered saline or fibronectin fragment, an osteoarthritis
   trigger. We identified 590 differentially spliced genes between
   conditions, with FN-f inducing splicing events similar to those in
   primary osteoarthritis tissue. CRISPR/Cas9 mimicking of an SNRNP70
   splicing event observed in osteoarthritis induced an
   osteoarthritis-like expression pattern. Integration with genotyping
   data revealed 7188 splicing quantitative trait loci (sQTL) affecting
   3056 genes, including 738 and 343 condition-specific sQTLs for resting
   and fibronectin fragment, respectively. Colocalization with
   osteoarthritis GWAS identified 6 putative risk genes. Our study
   highlights the significant impact of alternative splicing in
   osteoarthritis and provides potential therapeutic targets.
   Subject terms: RNA splicing, Cartilage
     __________________________________________________________________
   The authors identify thousands of genetic variants affecting RNA
   splicing in primary human chondrocytes and link several of them to
   osteoarthritis risk using genome editing and computational approaches.
Introduction
   Osteoarthritis (OA) is a chronic, debilitating joint disease affecting
   over 500 million people worldwide^[45]1. Despite its prevalence,
   effective treatments to prevent disease progression remain elusive, due
   in part to our limited understanding of the underlying molecular
   mechanisms^[46]2. Genetic factors play a substantial role in OA risk,
   with an estimated heritability of over 50%^[47]3. Genome-wide
   association studies (GWAS) have identified hundreds of loci associated
   with OA risk^[48]4,[49]5, yet understanding their functional impact
   remains challenging, as most associated single nucleotide polymorphisms
   (SNPs) reside in non-coding genomic regions^[50]6,[51]7.
   One powerful method to determine the mechanisms through which
   disease-associated variants act is to identify quantitative trait loci
   (QTL) for a feature of interest (e.g., gene expression). Once mapped,
   those QTLs can be colocalized with GWAS risk variants to identify
   putative causal mechanisms. We and others have colocalized expression
   QTL data sets with OA GWAS to identify putative causal
   genes^[52]4,[53]8,[54]9. Despite these successes, many OA GWAS loci
   remain unexplained. One possible reason is that the disease-associated
   variants might influence traits other than gene expression. This may be
   due to alternative splicing, as existing literature has described
   substantial contributions of both expression QTLs and splicing QTLs to
   disease risk; however, these disease-associated eQTLs and sQTLs often
   influence disease risk via different genes^[55]10–[56]12. While OA
   tissue is characterized by distinct changes in splicing
   patterns^[57]13; sQTLs have yet to be mapped in cartilage or other
   OA-relevant cell types.
   To address this critical need, we mapped splicing QTLs in primary human
   chondrocytes, the cell type responsible for maintaining the
   extracellular matrix of articular cartilage. Splicing QTLs were mapped
   in chondrocytes both in the resting state and in response to
   fibronectin fragment (FN-f), a known OA trigger. FN-fs are present in
   the cartilage and synovial fluid of OA patients^[58]14 and in cell
   culture studies have been shown to initiate a range of catabolic
   signaling pathways characteristic of chondrocytes isolated from OA
   tissue, mimicking the OA-like state^[59]15–[60]17. This approach allows
   us to capture the transition from a normal to an OA phenotype and
   identify variants that exert effects in both homeostatic and
   pathological settings.
   In this work, we analyzed RNA-seq data from 101 human chondrocyte
   samples treated with either PBS (resting) or FN-f, as well as 16 OA
   donor samples. We identified hundreds of genes that were differentially
   spliced between PBS-treated and either FN-f treated or OA chondrocytes.
   Differential splicing events induced by FN-f mimicked those seen in OA
   tissue, underscoring the accuracy of our FN-f model. We used CRISPR to
   mimic a splicing event observed in both FN-f treated normal
   chondrocytes and in OA chondrocytes and found that it induced OA-like
   expression changes. We identified 7188 splicing quantitative trait loci
   (sQTL) affecting 3056 genes, over 1000 of which exhibited
   condition-specific effects. Colocalization of these sQTLs with OA GWAS
   identified 6 putative OA risk genes, including putative candidate
   PBRM1. This approach demonstrates the potential of sQTL analysis to
   uncover genetic factors contributing to OA susceptibility. Our findings
   both improve the mechanistic understanding of the disease and identify
   genes that can be further studied and potentially targeted for
   therapeutic intervention.
Results
FN-f induces alternative splicing in human chondrocytes
   To determine how FN-f stimulation influences RNA splicing, we
   reanalyzed our existing RNA-seq data sets obtained from primary human
   chondrocytes isolated from 101 non-OA donors (donor information in
   Supplementary Data [61]1) treated with either PBS or FN-f, a known OA
   trigger. Initial quality control revealed minor batch effects
   associated with the RNA extraction kit and FN-f batch (Supplementary
   Fig. [62]1a). We corrected these technical confounders using the limma
   package^[63]18, resulting in improved sample clustering by biological
   condition (Supplementary Fig. [64]1b). Differential analysis using
   LeafCutter identified 974 differential splice junctions corresponding
   to 590 genes (adjusted p < 0.05, |∆PSI| > 0.15, Supplementary
   Data [65]2). Hierarchical clustering of differentially spliced intron
   junctions revealed distinct patterns between PBS resting and FN-f
   treated samples (Fig. [66]1a). Over ninety percent of these
   differential splicing events were previously annotated in GENCODE
   release 45 (GRCh38.p14) as reported by LeafCutter. (Fig. [67]1b).
   Furthermore, both our annotated and unannotated splice junctions
   exhibited strong enrichment for canonical splicing motifs
   (Supplementary Fig. [68]1c). Together, these results support the
   validity of our splicing analysis.
Fig. 1. FN-f induces alternative splicing events in human chondrocytes.
   [69]Fig. 1
   [70]Open in a new tab
   a Heatmap of differential splicing events between PBS and FN-f treated
   chondrocytes (n = 101 biologically independent non-OA donors), with
   each donor providing one chondrocyte culture split for PBS and FN-f
   treatment. Rows show PSI values for 621 intron junctions from
   LeafCutter clusters (|∆PSI| > 0.15, adjusted p < 0.05), derived from
   590 genes. Significance determined using two-sided
   Dirichlet-multinomial regression with FDR correction by LeafCutter.
   Ancestry: AMR (Ad-mixed American), AFR (African), EUR (European), SAS
   (South Asian). b Pie chart illustrates the distribution of differential
   alternative splicing intron junction (n = 974) types. c Protein-protein
   interaction network of 590 differentially spliced genes using
   high-confidence interactions from STRING database v12.0 (interaction
   score > 0.900). Network shows genes with protein-protein interactions
   above the confidence threshold. Colors highlight pathway groups, bold
   circles indicate genes also differentially spliced in OA tissue, text
   indicates genes differentially spliced in both FN-f and OA conditions.
   d Venn diagram showing overlap between FN-f-induced splicing changes
   (n = 101 donors) and OA-associated splicing changes (n = 16 OA donors
   versus n = 101 non-OA controls). Statistical significance determined
   using two-sided Dirichlet-multinomial regression with FDR correction by
   LeafCutter default method. e Box plots showing ∆PSI in FN-f response
   for introns differentially spliced in OA. Each dot represents an
   individual splice junction (n = 237 for increased PSI in OA, n = 221
   for decreased PSI in OA). The analysis compares OA vs PBS splicing
   changes (from 16 OA and 101 non-OA biological donors) with FN-f vs PBS
   changes (from the same 101 non-OA donors). Box plots show median
   (center line), 25th and 75th percentiles (box bounds), minimum and
   maximum values within 1.5× interquartile range (whiskers), with
   outliers as individual points. Statistical test: one-sided Wilcoxon
   signed-rank test against theoretical median of zero (p = 2.23 × 10^−8
   for increased PSI group; p = 1.23 × 10^−6 for decreased PSI group). No
   adjustments were made for multiple comparisons.
   We performed pathway and Gene Ontology (GO) enrichment analysis on the
   differentially spliced genes to understand the potential functional
   implications of these splicing changes. For FN-f-stimulated
   chondrocytes as compared to PBS controls, this analysis revealed
   significant enrichment in several pathways and processes, some of which
   had possible connections to OA pathogenesis (Supplementary
   Fig. [71]2a). Key enriched pathways included signal transduction, FGFR1
   signaling in disease, deleted in colorectal cancer-mediated attractive
   signaling, and osteoclast differentiation. Enriched GO terms included
   anatomical structure morphogenesis, actin cytoskeleton organization,
   and cell motility.
   To further interrogate the relationship between differentially spliced
   genes, we performed protein-protein interaction network analysis and
   detected 64 nodes with high-confidence interactions (Interaction
   score > 0.900). This network highlighted subnetworks corresponding to
   critical pathways such as GPCR signaling, signal transduction,
   osteoclast differentiation, hedgehog signaling, and steroid hormone
   biosynthesis (Fig. [72]1c). Taken together, these results demonstrate
   that FN-f induces large-scale changes in alternative splicing that
   converge on specific pathways and processes, many of which have
   possible relevance to OA.
FN-f induces OA-like RNA splicing in human chondrocytes
   To determine if our FN-f-treated chondrocytes exhibited similar
   splicing patterns to those found in OA tissue, we performed RNA-seq on
   chondrocytes isolated from cartilage obtained from 16 donors who
   underwent knee replacement surgery. Comparison between our PBS-treated
   non-OA chondrocytes and OA chondrocytes revealed 466 significantly
   differentially spliced intron junctions corresponding to 321 genes
   (adjusted p < 0.05, |∆PSI| > 0.15, Supplementary Fig. [73]2b, c and
   Supplementary Data [74]2). Genes that were differentially spliced in OA
   compared to non-OA chondrocytes (treated with PBS) were enriched in
   multiple KEGG pathways, including MCM complex, Rho GTPase cycle, and
   fatty acid metabolism (Supplementary Fig. [75]2d and Supplementary
   Data [76]3) and GO terms, including positive regulation of GTPase
   activity, cell morphogenesis, and regulation of filopodium assembly.
   Seventy-five intron junctions (corresponding to 75 distinct genes) were
   significantly altered in both FN-f-stimulated and OA chondrocytes
   (Fig. [77]1d, right panel). 75% of the splicing changes shared between
   OA and FN-f treated samples showed a concordant directional effect, and
   the effect sizes were significantly correlated (R^2 = 0.27, p < 0.001,
   linear model regression, Supplementary Fig. [78]2e). Several of these
   OA-associated differential spliced genes were also present in the
   protein-protein interaction networks generated from FN-f induced
   splicing events (Fig. [79]1c in bold), suggesting potential functional
   connections. This exploratory network analysis highlighted genes with
   established roles in cartilage biology or processes relevant to OA
   pathogenesis. For instance, SNRNP70 is a splicing factor that has been
   implicated in rheumatoid arthritis^[80]19; MALAT1 is a lncRNA involved
   in inflammation and chondrocyte proliferation^[81]20, and NCOR2 is a
   nuclear receptor co-repressor that is hypermethylated and downregulated
   in OA tissue^[82]21.
   Next, we sought to determine the extent to which splicing events in
   FN-f-treated chondrocytes mirror those seen in OA. Differential
   splicing events induced by FN-f were significantly correlated with
   those seen in OA tissue (linear regression, p < 0.001, Supplementary
   Fig. [83]2f). To investigate this further, we divided the 466 introns
   that were differentially spliced between OA and non-OA (PBS-treated)
   chondrocytes into those that showed increased or decreased percent
   spliced in (PSI) in OA samples. On average, introns that exhibited
   increased PSI in OA tissue also exhibited increased PSI in response to
   FN-f (Fig. [84]1e). Conversely, introns that exhibited decreased PSI in
   OA tissue also exhibited decreased PSI in response to FN-f
   (Fig. [85]1e). Our FN-f model mimics the early events in OA onset, and
   we wouldn’t expect it to fully recapitulate the changes seen in
   end-stage OA tissue; however, the agreement between splicing changes in
   OA and FN-f-treated chondrocytes, in terms of both broad directional
   changes and specific high-confidence alterations, suggests that the
   FN-f-treated cells provide a valuable model to interrogate OA-related
   alternative splicing events. The identified set of consistently altered
   splicing events provides a focused group of targets for future studies
   aimed at understanding and potentially intervening in OA progression.
   Splicing events that differed between FN-f-treated and OA samples were
   enriched for GO terms, including ion transport processes, specific
   protein modifications, and metabolic pathways (Supplementary
   Data [86]3), which provide insight into which aspects of OA are and are
   not fully modeled via FN-f treatment.
Forced splicing of SNRNP70 induces OA-like gene expression
   We identified SNRNP70, a critical component of the U1 snRNP complex, as
   differentially spliced in both FN-f-stimulated and OA chondrocytes
   compared to PBS (resting). RNA-seq analysis revealed significant
   changes in SNRNP70 splicing, particularly between exons 7 and 8
   (Fig. [87]2a). The alternative exon 8 (chr19:49,102,114-49,103,587)
   showed reduced inclusion in FN-f and OA conditions, with ∆PSI
   differences of 22% (FN-f/PBS) and 23% (OA/PBS). This indicates
   alternative exon 8 skipping in 67% and 68% of transcripts in FN-f and
   OA conditions, respectively, compared to 45% in PBS (resting).
   Exon-level expression analysis confirmed significant downregulation
   (p < 0.01) of the alternative exon 8 in both conditions (Fig. [88]2b).
   This differential splicing event was confirmed via qPCR (Supplementary
   Fig. [89]3). Western blotting for SNRNP70 revealed a single band
   representing the expected protein size, with reduced protein levels
   under FN-f conditions matching the transcriptomic data (Supplementary
   Fig. [90]4).
Fig. 2. Deletion of SNRNP70 alternative exon 8 mimics OA-related splicing
patterns.
   [91]Fig. 2
   [92]Open in a new tab
   a RNA-seq signal track for PBS-treated, FN-f-treated, and OA
   chondrocytes at the SNRNP70 gene locus. Below the signal tracks are
   transcript annotations representing the MANE Select transcript
   (ENST00000598441.6) and a transcript containing the alternative exon 8
   (ENST00000401730.5). b Barplot represents log[2] fold change for
   expressed exons with FN-f and OA compared to PBS. c Schematic of
   CRISPR/Cas9 genome editing strategy for SNRNP70 alternative exon 8
   deletion. d Barplot comparing log[2] fold change for expressed exons
   after SNRNP70 alt ex8 deletion. e Barplot of KEGG and Reactome pathways
   and GO terms enriched in differentially expressed genes between
   alternative exon 8 (alt ex 8) heterozygous knockdown (Het-KD) and
   wildtype (WT). Enrichment was calculated using cumulative
   hypergeometric distribution of overlap genes to the total background
   gene set, with statistical significance indicated by -log[10] of the
   p-value. One-sided hypergeometric tests were used to identify
   significant enrichment, without adjustment for multiple comparisons. A
   total of 135 differentially expressed genes (padj <0.01,
   |Log[2]FoldChange| > 2) were analyzed, with 71 upregulated and 64
   downregulated genes in Het-KD compared to WT. f Box plots showing gene
   expression changes (log[2]FoldChange) in FN-f vs PBS for genes that
   were differentially expressed in the SNRNP70 alternative exon 8
   heterozygous knockdown (Het-KD) vs wildtype (WT) comparison. The
   analysis examines whether genes upregulated in Het-KD/WT also tend to
   be upregulated in FN-f/PBS, and whether genes downregulated in
   Het-KD/WT tend to be downregulated in FN-f/PBS. Each dot represents an
   individual gene (n = 60 for genes upregulated in Het-KD vs WT, n = 57
   for genes downregulated in Het-KD vs WT). Box plots show median (center
   line), 25th and 75th percentiles (box bounds), 5th and 95th percentiles
   (whiskers), and outliers as individual points. Statistical analysis:
   one-sided Wilcoxon signed rank test with continuity correction
   (upregulated genes: p-value = 0.008283; downregulated genes:
   p-value = 0.01869). No adjustments were made for multiple comparisons.
   To assess the functional role of SNRNP70 alternative exon 8, we
   employed CRISPR-Cas9 to delete this region in primary human
   chondrocytes (Fig. [93]2c). Our deletion strategy targeted guide RNAs
   to both sides of a 1735 bp region encompassing the alternative exon 8
   (Supplementary Fig. [94]5a). PCR screening of 46 single-cell-derived
   colonies revealed 26% with heterozygous and 0% with homozygous deletion
   of the alternative exon 8 (Supplementary Fig. [95]5b, c). Sanger
   sequencing of Het-KD colonies confirmed the deletion, with an
   additional cytosine deletion beyond the expected cut site
   (Supplementary Fig. [96]5d).
   To determine the phenotypic impact of heterozygous exon 8 deletion, we
   performed RNA-seq on both edited and unedited colony-expanded cells.
   The differential analysis confirmed decreased SNRNP70 alternative exon
   8 inclusion (Fig. [97]2c) and identified 135 differentially expressed
   genes (DESeq2, FDR-adjusted p-value < 0.01). Genes that were
   upregulated in Het-KD cells were enriched for multiple GO terms and
   KEGG pathways, including IL-17 signaling, chemokine receptor binding,
   and rheumatoid arthritis (Fig. [98]2e and Supplementary Data [99]4).
   Genes that were downregulated in Het-KD cells were enriched for
   pathways including L1 and Ankyrin interactions and O-glycan
   biosynthesis (Fig. [100]2e and Supplementary Data [101]4). Differential
   splicing analysis revealed 259 differential splicing events (233
   differential splicing genes) in Het-KD cells, although these changes
   did not clearly correlate with those observed in FN-f treated cells
   (Supplementary Fig. [102]5e).
   Interestingly, the directionality of gene expression changes in Het-KD
   cells significantly correlated with those in FN-f-stimulated
   chondrocytes (p < 0.05 for both up- and downregulated genes,
   Fig. [103]2f). Genes that were upregulated both in response to FN-f and
   in response to alternative exon 8 deletion included CXCL8, CXCL2, IL11,
   MMP12, and MMP13, which are associated with inflammation and
   extracellular matrix degradation in OA^[104]22–[105]24. Downregulated
   genes included PEG3, FGL2, SEMA3E, CSRNP3, and HLF, which have been
   implicated in chondrocyte homeostasis^[106]25–[107]29. These results
   suggest that the exclusion of SNRNP70 alternative exon 8 in response to
   cartilage matrix damage modeled by FN-f treatment may be responsible
   for some of the other expression changes associated with the OA
   phenotype.
Genetic variants impact splicing in chondrocytes
   We performed splicing QTL (sQTL) analysis on both PBS (resting) and
   FN-f stimulated chondrocyte samples to determine the impact of genetic
   differences on chondrocyte mRNA splicing. We tested the association of
   each RNA splicing event with genetic variants within ±100 kb of the
   start and end points of the splice intron junctions, using PSI values
   adjusted for confounding factors (Supplementary Fig. [108]6a). We
   identified a total of 5873 unique sQTLs corresponding to 2575 sGenes
   (genes with at least one significant sQTL) in PBS resting, and 4606
   sQTLs corresponding to 2132 sGenes in FN-f stimulated chondrocytes
   (QTLtools^[109]30, FDR < 0.05, Fig. [110]3a and Supplementary
   Data [111]5). Over half (3400 in PBS and 2631 in FN-f) of sQTLs were
   not previously detected in any GTEx tissue (Supplementary
   Fig. [112]6e). We attempted to perform sQTL analysis on our 16 OA donor
   samples, but due to the very small sample size, we identified only a
   single sQTL.
Fig. 3. Condition-specific sQTL discovery in resting and activated
chondrocytes.
   [113]Fig. 3
   [114]Open in a new tab
   a Venn diagram showing the overlap between sGenes identified in PBS and
   FN-f treated conditions. Linear mixed-effect modeling allowed the
   detection of 677 condition-specific sGenes (477 PBS-specific and 200
   FN-f-specific). b The distribution of distances between sSNPs and the
   affected splice intron junctions (kb). c Percentage of sSNPs located
   within the same gene (within a gene) or outside the gene. d–f Examples
   of d shared (RHBDF2), e PBS-specific (SLC26A4), and f FN-f-specific
   (MAPK8) sQTLs. Each panel contains locus zoom plots showing the
   -log[10](p-value) of association between SNPs and splicing across
   genomic positions for both PBS and FN-f conditions. Gene structures are
   displayed beneath each locus zoom plot. Each panel also includes box
   plots displaying the relationship between genotype and inverse normal
   transformed PSI values for the indicated splice junction in both PBS
   and FN-f conditions (n = 101 independent donor samples per condition).
   Boxes indicate the interquartile range (25th–75th percentile) with a
   line at the median (50th percentile). Whiskers extend to the minimum
   and maximum values within 1.5× the interquartile range, with outliers
   shown as individual points. The figure shows the lead SNP ID, effect
   size (Beta), and p-value for each condition. Shared sQTLs show
   significant associations in both PBS and FN-f conditions. PBS-specific
   sQTLs show significant association only in PBS but not in FN-f
   condition. FN-f-specific sQTLs show significant association only in
   FN-f but not in PBS condition. Nominal p-values were calculated using
   QTLtools, which performs two-sided tests for association by converting
   correlation coefficients to p-values using a t-distribution of freedom
   (n-2, where n = 101 samples).
   Conditional analysis revealed 182 secondary signals, 13 tertiary
   signals, and 2 quaternary signals for PBS-treated chondrocytes and 121
   secondary signals and 1 tertiary signal for FN-f-treated chondrocytes
   (Supplementary Fig. [115]6b). In total, 116 PBS sGenes and 82 FN-f
   sGenes had two or more independent signals. For example, the CAST gene
   exhibited multiple independent signals associated with the splice
   intron junction chr5:96768431–96776404 in PBS resting, while the MICA
   gene exhibited multiple independent signals with the splice intron
   junction chr6:31410797–31411209 in FN-f stimulated chondrocytes
   (Supplementary Fig. [116]7). To further characterize the genetic
   architecture of splicing regulation, we systematically assessed allelic
   heterogeneity across genes with multiple sQTLs. By examining linkage
   disequilibrium (LD) patterns (R^2 < 0.7) between sQTLs in the same
   gene, we identified 760 genes in PBS tissue (29.5% of 2575 sGenes) and
   576 genes in FN-f tissue (27.0% of 2132 sGenes) in which two or more
   independent variants influenced splicing in the same gene.
   Additionally, we investigated the relationship between splicing
   regulation and gene expression by analyzing the overlap between our
   sQTLs and published eQTLs^[117]9 (Supplementary Data [118]5). We found
   that the vast majority of our sQTLs (5557 in PBS and 4302 in FN-f)
   showed no overlap (R^2 < 0.7) with eQTLs affecting the same gene,
   suggesting that the genetic variants influencing splicing at these loci
   did not also influence expression levels. While a smaller subset of
   sQTLs (513 in PBS and 426 in FN-f) did overlap with eQTLs. Retesting
   for sQTLs while conditioning on the genotype of the lead eQTL SNP
   suggests that over 99% of these sQTLs that were in high LD with an eQTL
   (512 out of 513 in PBS and 424 out of 426 in FN-f) also induced changes
   in gene expression. In summary, approximately 10% of our detected sQTLs
   also influenced gene expression.
   We analyzed the genomic distribution of sQTLs (sSNPs) relative to their
   associated splice intron junctions. Lead sSNPs were largely close to
   splice junctions, with 43% and 45% of lead SNPs within 5 kb of the
   affected splice junctions for PBS and FN-f, respectively
   (Fig. [119]3b). Further analysis revealed that the majority of sSNPs
   were located within the same gene as their associated splice intron
   junction (Fig. [120]3c). Taken together, these results both support the
   validity of our sQTLs and suggest that sSNPs largely act at the
   affected splice junctions rather than via distal mechanisms.
   To determine if any of the effects of variants were dependent on the
   condition, we used a linear mixed-effect model and identified 1635
   sGenes that exhibited significant genetic and condition interaction
   effects (lme4^[121]31, ANOVA, p < 0.05). To compile a high-confidence,
   large-effect size list of condition-specific sQTLs, we further refined
   this list to include only those that were identified as an sQTL in only
   one condition and exhibited an absolute beta difference between
   conditions of greater than or equal to 0.2. In total, we identified 677
   high-confidence condition-specific sGenes (Fig. [122]3a). 400 lead SNPs
   had a stronger genetic effect on splicing in PBS-treated cells, and 277
   lead SNPs had a stronger genetic effect on splicing in FN-f-treated
   cells. Examples of shared, PBS-specific, and FN-f-specific sQTLs are
   depicted in Fig. [123]3d–f. Several of these condition-specific sGenes
   have known roles in cartilage biology and/or OA. For example,
   Fig. [124]3e shows a PBS-specific condition sQTL for SLC26A4 involving
   the splice junction chr7:107694480-107694621. SLC26A4, an anion
   transporter, may influence chondrocyte metabolism in homeostatic
   conditions, potentially affecting cartilage maintenance^[125]32.
   Fig. [126]3f demonstrates an FN-f-specific sQTL for MAPK8, associated
   with the splice junction chr10:48363284-48401612. MAPK8 (JNK1) is
   involved in inflammatory signaling, and JNK inhibition is currently
   being investigated as a treatment for OA^[127]33,[128]34. However,
   knockout of Jnk1 or Jnk2 increased the senescence burden and the extent
   of age-related osteoarthritis in mice^[129]35. The complex role of JNK
   signaling in OA may result from high activity in a rare sub-population
   of chondrocytes that serves an “inflammation-amplifying” role in OA
   cartilage^[130]36.
RNA-binding proteins reveal putative sQTL mechanisms
   RNA-binding proteins (RBPs) play crucial roles in post-transcriptional
   regulation, influencing various cellular processes, including
   inflammation and metabolic changes related to OA^[131]37,[132]38.
   Recent studies have highlighted the dysregulation of RBPs in
   chondrocytes, synovial fibroblasts, and osteoblasts, indicating their
   significant impact on OA progression^[133]39,[134]40. To gain insight
   into the putative mechanisms through which our sQTLs were influencing
   splicing in chondrocytes, we leveraged publicly available enhanced
   crosslinking and immunoprecipitation (eCLIP) data from the ENCODE
   database^[135]41. This dataset encompasses binding sites for 140 RBPs
   across three cell types: HepG2 (n = 85), K562 (n = 118), and SM-9MVZL
   (n = 2). While these cell lines are not OA-specific, similar data is
   not currently available in chondrocytes, and presumably, many of these
   binding sites are shared across cell types.
   Using the QTLtools fenrich module, we assessed the statistical
   significance of RBP enrichment through permutation tests. For PBS
   (resting) lead sQTLs, we identified 12 significantly enriched RBPs
   (odds ratio > 1, empirical p < 0.05), including SF3B4, EFTUD2, and
   PRPF8. The FN-f condition yielded 7 enriched RBPs, including UP3,
   EFTUD2, and SF3B4 (Fig. [136]4a and Supplementary Data [137]6). We
   found 255 sQTL lead variants that were located within the binding site
   of at least one RNA-binding protein (RBP) and showed significant
   correlation between the splicing pattern and RBP gene expression
   (Pearson’s R^2 > 0.15, p < 0.01; Supplementary Data [138]6).
Fig. 4. RNA-Binding protein associations with sQTLs.
   [139]Fig. 4
   [140]Open in a new tab
   a Volcano plot depicting the effect size (x-axis) and -log[10] p-values
   (y-axis) of RBP sites enrichment in sQTLs identified in PBS- or
   FN-f-treated chondrocytes. P-values and odds ratios were calculated
   using QTLtools Fenrich module with 1000 permutations (“Methods”). The
   enrichment analysis uses a two-sided empirical permutation test to
   determine whether RBP binding sites are over- or under-represented in
   regions containing sQTLs compared to control regions, without
   adjustment for multiple comparisons. Data points are shown for
   p-value < 0.05 and odds ratios > 1. b Analysis of AATF RBP binding and
   expression correlation with the rs11871958 sQTL in the SNHG29 gene.
   Locus zoom plots show -log[10](p-value) of SNP-splicing associations
   across the genomic region in PBS and FN-f conditions (PBS:
   p = 1.48 × 10^−21, β = −0.9; FN-f: p = 1.06 × 10^−24, β = −0.99, minor
   allele frequency = 0.1832). These association statistics were
   calculated using QTLtools. The middle track displays AATF eCLIP signal
   (RPM) with the sQTL SNP location highlighted. Box plots show normalized
   PSI values by rs11871958 genotype (n = 202 samples from 101
   biologically independent donors, each providing PBS and FN-f samples).
   Box plots display median (center line), 25th–75th percentiles (box),
   minimum-maximum values within 1.5× interquartile range (whiskers), and
   outliers (points). The scatter plot shows correlation between AATF gene
   expression (VST-normalized DESeq values) and normalized PSI values
   across all samples (R^2 = 0.408, p = 2.22 × 10^−28). Pearson
   correlation was used with a two-sided linear regression model to
   calculate the p-value, without adjustment for multiple comparisons. c
   Analysis of EFTUD2 RBP binding and expression correlation with the
   rs6809764 sQTL in the PCYT1A gene, following the same format as panel
   b. QTLtools statistics for this sQTL: PBS (p = 4.65 × 10^−8, β = 0.44);
   FN-f (p = 2.29 × 10^−7, β = 0.522, minor allele frequency = 0.3911).
   The correlation between EFTUD2 gene expression and normalized PSI
   values (n = 202 samples) shows R^2 = 0.164, p = 7.84 × 10^−9,
   calculated using the same statistical methods.
   An example of a putative RBP-mediated sQTL is rs11871958, which is
   associated with alternative splicing of SNHG29 and overlaps an AATF RBP
   binding site (Fig. [141]4b). The minor C allele of rs11871958 is
   associated with decreased PSI of the associated intron junction
   (chr17:16439414-16439528). Interestingly, AATF gene expression is
   strongly correlated with the PSI of the associated intron junction,
   lending further support to the role of AATF in this splicing event
   (R^2 = 0.458, p = 2.233 × 10^−28). SNHG29, a long non-coding RNA, has
   been previously reported to regulate the miR-223-3p/CTNND1 axis and
   influence the Wnt/β-catenin signaling^[142]42, which has been
   previously implicated in OA pathogenesis^[143]43.
   Another example is rs6809764, which was identified as an sQTL for
   PCYT1A and overlaps an EFTUD2 binding site (Fig. [144]4c). The minor G
   allele of rs6809764 is associated with increased PSI of the intron
   junction (chr3:196270541-196287615). Gene expression of EFTUD2
   correlated with decreased junction usage (R^2 = 0.154,
   p = 7.394 × 10^−9), which supports EFTUD2 as a possible mediator of
   this sQTL. PCYT1A encodes an enzyme crucial for phosphatidylcholine
   biosynthesis^[145]44. It has been observed to localize to the nuclear
   envelope in hypertrophic chondrocytes of bone growth plates, coinciding
   with a significant increase in cell volume^[146]44. This localization
   suggests a potential role for PCYT1A in chondrocyte morphology changes
   during hypertrophy, a process relevant to cartilage development and OA
   pathogenesis^[147]45.
   These findings provide insights into the regulatory mechanisms of
   chondrocyte splicing events. However, we acknowledge the limitations of
   using non-chondrocyte cell lines for RBP binding data. Further
   biochemical studies in relevant cell types will be crucial to fully
   elucidate the mechanisms through which these genetic variants alter
   splicing patterns in cartilage tissue.
Colocalization analysis reveals putative OA risk genes
   To understand the potential role of splicing in OA risk, we performed
   colocalization analysis between our sQTLs and 100 independent OA GWAS
   loci spanning 11 OA-related phenotypes reported by Boer et al.^[148]4.
   We selected sQTLs whose lead SNP was in moderate to high LD (r^2 > 0.5)
   with a lead SNP from any of the 11 OA GWAS data SNPs and tested for
   colocalization using the coloc package (Supplementary Data [149]7). In
   total, we identified 6 colocalized signals corresponding to 6 unique
   sGenes across 4 OA phenotype subtypes (Table [150]1 and Fig. [151]5).
Table 1.
   Chondrocyte sQTLs that colocalize with OA GWAS loci
   Gene Intron Junction rsID Position Minor Allele MAF PPH3 PPH4 (PPH4)/
   (PPH3 + PPH4) GWAS phenotype rsID (GWAS) sQTL Condition GTEX Tissue
   Count
   RNF144B chr6:18399699–18401957 rs1886248 chr6:18399163:C:T C 0.4059
   0.00267 0.997 0.997 FingerOA rs9396861 Both 14
   WWP2 chr16:69925484–69929448 rs7192245 chr16:69921902:T:C T 0.2228
   0.179 0.714 0.780 KneeOA rs34195470 PBS 0
   COLGALT2 chr1:183978520–184037581 rs74767794 chr1:184036994:A:G G 0.297
   0.0302 0.921 0.9683 THR rs1327123 PBS 0
   GCAT chr22:37808163–37809949 rs13057823 chr22:37792277:C:A A 0.223
   0.0598 0.924 0.9392 THR rs12160491 Both 47
   HMGN1 chr21:39347292–39347379 rs2249666 chr21:39347315:C:T T 0.451
   0.0994 0.786 0.8877 TJR rs9981884 PBS 31
   PBRM1 chr3:52658315–52662133 rs7628578 chr3:52594305:A:T T 0.4653 0.270
   0.717 0.7263 THR rs3774354 Fn-f 10
   [152]Open in a new tab
   The table shows significant colocalization results between sQTL and
   GWAS signals (PP H4 > 0.7) with the most significant splice intron
   junction. The number of GTEx tissues (out of 50) where the same sQTL is
   detected is also indicated. PP posterior probability, FingerOA
   osteoarthritis of the finger, KneeOA osteoarthritis of the knee, THR
   total hip replacement.
Fig. 5. Six sQTLs colocalize with OA GWAS signals.
   [153]Fig. 5
   [154]Open in a new tab
   a Total hip replacement (THR) GWAS signal colocalizes with an sQTL
   found in both conditions for PBRM1 associated with splice intron
   junction chr3:52658315–52662133. b THR GWAS signal colocalizes with an
   sQTL found in both conditions for GCAT associated with splice intron
   junction chr22:37808163–37809949. c Total joint replacement (TJR) GWAS
   signal colocalizes with an sQTL found in both PBS-treated chondrocytes
   for HMGN1 associated with splice intron junction
   chr21:39347292–39347379. d Finger OA GWAS signal colocalizes with an
   sQTL found in both conditions for RNF144B associated with splice intron
   junction chr6:18399699–18401957. e Knee OA GWAS signal colocalizes with
   an sQTL found in PBS-treated chondrocytes for WWP2 associated with
   splice intron junction chr16:69925484–69929448. f THR GWAS signal
   colocalizes with an sQTL found in PBS-treated for COLGALT2 associated
   with splice intron junction chr1:183978520–184037581.
   To the best of our knowledge, these splicing events have not previously
   been colocalized with OA GWAS signals; however, three of these genes
   have been implicated in OA via colocalization with expression QTLs.
   GCAT and HMGN1 have been previously associated with knee and/or hip OA
   via colocalization with an osteoclast-specific eQTL^[155]46. RNF144B
   was previously linked to OA via colocalization between OA GWAS variants
   and eQTLs from chondrocytes, tibial nerves, and testis^[156]4,[157]9.
   Our results build upon these previous studies by providing more
   mechanistic detail into how these variants act, presumably via their
   impact on alternative splicing.
   Three of our colocalized genes have not been previously linked to OA
   via colocalized eQTLs but do have supporting evidence for a role in OA.
   WWP2 and COLGALT2 have been implicated in OA by their proximity to
   mQTLs that colocalize with OA GWAS variants and control methylation
   sites near WWP2 and COGALT2^[158]47. Methylation at these loci was
   confirmed to affect WWP2 and COGALT2 via epigenome
   editing^[159]48,[160]49. To the best of our knowledge, one of our
   colocalized sGenes (PBRM1) has not been previously linked to OA via
   genetic methods but was found to be hypomethylated in OA
   cartilage^[161]50,[162]51.
   These findings highlight the potential role of alternative splicing in
   mediating OA risk and demonstrate the value of integrating sQTL data
   with GWAS results to uncover the potential mechanisms underlying
   genetic associations.
Discussion
   This study establishes a comprehensive sQTL map in primary
   chondrocytes, providing insights into the relationship between genetic
   variation, alternative splicing, and OA risk. By integrating sQTL
   analysis with OA GWAS data and functional validation, we have
   identified potential mechanisms underlying OA pathogenesis and
   highlighted putative OA risk genes for further investigation and future
   therapeutic development.
   The splicing alteration observed in both FN-f stimulated and OA
   chondrocytes highlights the role of alternative splicing in OA
   progression. Our identification of 974 and 466 differentially spliced
   intron junctions in FN-f-stimulated and OA chondrocytes, respectively,
   suggests broad splicing dysregulation in OA. The overlap between
   FN-f-induced and OA-associated splicing events supports the relevance
   of our ex vivo model and suggests that some of these splicing changes
   may contribute to OA pathogenesis. However, while the overlap between
   FN-f treated cells and cells from OA tissue is significant, their
   correlation is modest, which may be due to the small number of OA
   samples available for this study. In addition, the OA tissue is
   obtained from patients undergoing knee replacement, which reflects the
   end-stage of the disease, while the FN-f stimulus of normal
   chondrocytes may induce splicing events important at earlier stages of
   OA. Therefore, while FN-f treatment is a controllable OA-relevant
   stimulus that provides insights into the genetics of OA risk, it does
   not fully recapitulate the splicing changes observed in end-stage OA.
   Our functional interrogation of SNRNP70 alternative splicing provides
   evidence for the impact of splicing changes on OA-related gene
   expression. The heterozygous deletion of SNRNP70 alternative exon 8 was
   a mimic of the reduced retention of alternative exon 8 that occurs
   during OA and in response to FN-f. While deletion of Snrnp70 itself is
   embryonic lethal in mice^[163]52, it was anticipated that chondrocytes
   could tolerate full deletion of alternative exon 8. The lack of
   colonies with homozygous alternative exon 8 deletions may represent a
   growth disadvantage as compared to intact or heterozygous deletion. The
   exact mechanism by which this splicing event alters chondrocyte biology
   is not entirely clear, but our data and the literature highlight
   several possibilities. Alternative exon 8 is located in the RNA
   recognition motif domain, which is critical for the protein RNA binding
   function, and its inclusion could alter the structure or function of
   the resulting protein. Alternatively, this splicing event could
   influence overall SNRNP70 expression levels, as transcripts that
   include alternative exon 8 are predicted by GENCODE v45 to induce
   nonsense-mediated decay (NMD)^[164]53. While the strongest impact of
   our CRISPR experiments was a sharp decrease in alternative exon 8
   usage, we also observed small but significant decreases in the usage of
   exons 9 and 10. This pattern is consistent with potential effects on
   transcript stability or NMD processes following exon 8 deletion, though
   the precise mechanism linking the deletion of alternative exon 8 to
   reduced usage of downstream exons requires further investigation.
   Alternative exon 8 deletion led to alterations in genes and pathways
   crucial for OA progression, including inflammatory signaling and matrix
   degradation, and suggests that alternative spliced genes may influence
   cellular phenotypes relevant to OA. The identification of other
   differentially spliced genes in both FN-f stimulated and OA
   chondrocytes, such as STS, ANKRD36C, and PIK3CD, provides potential
   targets for further investigation in OA. Interestingly, while the
   strongest impact of our CRISPR experiments was a sharp decrease in
   alternative exon 8 usage, we also observed small but significant
   decreases in exon 9 and exon 10 usage. The direct impact of alternative
   8 removal vs its downstream effect on exons 9 and 10 will require
   further investigation.
   Our sQTL analysis identified genetic variants influencing splicing in
   both resting and activated chondrocytes. The prevalence of local
   regulatory effects suggests that cis-acting splicing regulation
   contributes to how genetic variation influences OA risk. We identified
   hundreds of sQTLs that are influenced by condition (i.e., PBS vs FN-f),
   highlighting the need to conduct QTL studies in the correct biological
   context. What drives these condition-specific differences is
   challenging to determine without further experiments, but could be
   explained by differences in expression, splicing, nuclear localization,
   or post-translational modification of the splicing factors that mediate
   these events. Our RBP binding analysis identified potential regulatory
   proteins that may mediate these sQTL effects, with 137 of 140 analyzed
   RBPs showing detectable expression in chondrocytes. We identified 12
   and 7 significantly enriched RBPs in PBS and FN-f conditions,
   respectively, including key spliceosome components such as SF3B4,
   EFTUD2, and PRPF8. The examples of AATF-mediated regulation of SNHG29
   splicing and EFTUD2 potential role in PCYT1A splicing provide insights
   into putative mechanisms underlying our sQTLs. While these associations
   require experimental validation, and our analysis was limited by
   reliance on non-chondrocyte eCLIP data, these findings illuminate the
   complex regulatory networks that may drive chondrocyte splicing changes
   and provide promising candidate genes for future mechanistic studies.
   The colocalization analysis between our sQTLs and OA GWAS signals
   identified 6 colocalized signals. Three of these colocalized genes have
   been previously linked to OA via colocalized eQTLs, but our results
   build on this knowledge by suggesting that their pathogenic mechanism
   may act via alterations to mRNA splicing. While WWP2 has been
   functionally linked to OA through knockout studies, both WWP2 and
   COLGALT2 are implicated in OA by their proximity to colocalized mQTLs
   and have been shown to exhibit allelic expression imbalance in
   cartilage transcriptome studies, despite the absence of direct eQTL
   associations^[165]47,[166]54.
   One of our colocalized genes, PBRM1, provides a new potential OA
   mediator that has not previously been implicated by hypothesis-driven
   studies or genomic analysis. The rs3774354 GWAS signal was initially
   assigned to ITIH1 based on genomic proximity^[167]4. Earlier studies of
   the rs6976/rs11177 signal at the same locus implicated different genes,
   including GNL3 and SPCS1^[168]55, which showed allelic expression
   imbalance correlating with that OA risk genotype, while regulatory
   activity was demonstrated for SNPs near both ITIH1 and GLT8D1^[169]56.
   However, our colocalized sQTL suggests that altered splicing of PBRM1,
   rather than effects on ITIH1, may be the mechanism driving OA risk at
   this locus. As a SWI/SNF chromatin-remodeling complex
   component^[170]57, PBRM1 influences gene expression patterns. Its role
   in mesenchymal stromal cell differentiation is particularly intriguing
   in the context of OA. Previous research has shown that PBRM1 affects
   osteogenic differentiation and modulates SOX9 expression, a key
   regulator of chondrogenesis and expression of chondrocyte matrix
   genes^[171]58,[172]59. These observations suggest that PBRM1 might be
   involved in maintaining the delicate balance between different cell
   fates in joint tissues, a process often disrupted in OA.
   Interestingly, GCAT and HMGN1 have been associated with hipOA, knee
   and/or hip OA and/or all OA in osteoclast-specific eQTL^[173]46. Our
   findings extend this by providing evidence for their involvement in
   chondrocyte splicing regulation. GCAT, encoding glycine
   C-acetyltransferase, plays a crucial role in glycine metabolism in
   mitochondria^[174]60. Previous studies have shown that epigenetic
   downregulation of GCAT can lead to respiratory defects in aged
   fibroblasts, and glycine supplementation has been found to restore
   these defects^[175]60. While the direct link between GCAT and OA
   pathogenesis remains to be fully elucidated, our identification of GCAT
   through sQTL analysis suggests that alterations in mitochondrial
   function and glycine metabolism may contribute to OA risk, possibly
   through effects on cellular energy production or extracellular matrix
   synthesis. HMGN1 is a known non-histone chromatin remodeler and has
   been implicated in the regulation of chondrogenesis^[176]61. Furusawa
   et al. demonstrated that HMGN1 directly regulated the expression of
   SOX9. This suggests that alternative splicing may modulate HMGN1
   function and regulate chondrocyte-specific gene expression.
   RNF144B showed different patterns of condition specificity in our sQTL
   analysis compared to previous eQTL studies. We observed strong RNF144B
   sQTLs in both PBS- and FN-f-treated conditions, whereas the eQTL for
   RNF144B was highly specific to the FN-f-treated condition^[177]9. This
   discrepancy highlights how sQTL and eQTL analyses can provide
   complementary information for mRNA production and processing. Given
   RNF144B role in regulating DNA damage-induced apoptosis, its
   differential splicing could affect chondrocyte survival and turnover in
   OA.
   WWP2 has been previously implicated in OA through methylation studies.
   In particular, Roberts et al. identified a differentially methylated
   region (DMR) associated with the OA risk-conferring allele of
   rs34195470^[178]48. This DMR was found to act as a
   methylation-sensitive transcriptional repressor, with increased DNA
   methylation resulting in increased WWP2 expression, particularly of the
   WWP2-FL and WWP2-N transcripts. Our colocalization of a WWP2 sQTL with
   an OA GWAS signal provides a direct connection to OA and adds another
   layer to its regulatory complexity, potentially indicating that OA risk
   variants alter splicing via changes to DNA methylation. Further work is
   necessary to explore this hypothesis.
   COLGALT2 has also been previously associated with OA
   risk^[179]49,[180]62. Previous research has identified that the OA risk
   allele of rs11583641 corresponds to reduced DNA methylation at nearby
   CpGs and increased COLGALT2 expression^[181]63. Our sQTL findings now
   implicate alternative splicing as another potential mechanism by which
   COLGALT2 variation may influence OA risk, further highlighting the
   complex regulatory landscape of this gene in OA pathogenesis.
   While our study represents a significant advance in understanding the
   role of splicing in OA, it has several limitations that should be
   addressed in future research. Our sample size of 101 donors, while
   substantial for an sQTL study in chondrocytes, may not have captured
   all splicing-related OA GWAS genes. Future studies with larger sample
   sizes and expanded diversity will be crucial to fully elucidate the
   splicing landscape in OA. Additionally, while chondrocytes are highly
   relevant to OA, investigating splicing changes in other cell types
   involved in joint biology will provide a more comprehensive
   understanding of OA pathogenesis. Our FN-f stimulation model, while
   valuable, cannot fully recapitulate the complex joint environment.
   Advanced 3D culture systems, such as cartilage organoids or
   joint-on-a-chip models, could provide more physiologically relevant
   contexts for validating our findings. Furthermore, our study focused on
   a single time point of FN-f stimulation and collection of end-stage OA
   cartilage. Future studies considering other time-points or from
   patients at different stages of OA progression may capture splicing
   changes do not present here. While our functional validation of SNRNP70
   provides valuable insights, further experimental work is needed to
   understand the detailed functional consequences of the colocalized
   sQTLs we identified. Potential approaches for these investigations
   could include CRISPR-Cas9 mediated modulation of splice sites, minigene
   assays to confirm splicing effects, and functional assays to assess the
   impact on chondrocyte biology and OA-related processes.
   Despite these limitations, our study advances understanding of the role
   of alternative splicing in OA. We present an sQTL analysis in
   chondrocytes, and we built upon this resource by performing
   colocalization analysis and functional validation. The identification
   of these risk genes and splicing events provides potential targets for
   therapeutic intervention and biomarker development in OA. As we
   continue to unravel the complex genetic and molecular underpinnings of
   OA, integrative approaches that consider multiple layers of gene
   regulation will be instrumental in developing more effective diagnostic
   tools and targeted therapies for this debilitating disease.
Methods
Ethical statement
   Non-OA tissues were obtained from deceased donors with no known history
   of arthritis from the Gift of Hope Organ and Tissue Donor Network
   through the Rush Medical Center, Chicago, IL. OA tissues were obtained
   from patients undergoing knee joint replacement surgery at the
   University of North Carolina Hospitals. The University of North
   Carolina Institutional Review Board (IRB) reviewed the study protocol
   and determined that research involving de-identified tissues from
   deceased donors and surgical waste material does not constitute human
   subjects research under federal regulations, and therefore did not
   require IRB approval (Study Number 14-0189). Therefore, the requirement
   for written informed consent for the use of tissues in this study was
   waived for both OA and non-OA tissues. The families of deceased tissue
   donors do consent that tissue can be used for research purposes. All
   samples were de-identified, and no individually identifiable protected
   health information was accessible to investigators. Essential
   demographic information, including sex and age ranges can be found in
   Supplementary Data [182]1.
Dataset and sample information
   Primary human chondrocytes were isolated from talar cartilage tissue
   obtained from 101 non-osteoarthritis (non-OA) donors (101 PBS resting
   and 101 FN-f stimulated) and from 16 donors undergoing knee joint
   replacement due to OA. The complete dataset is available in the
   Database of Genotypes and Phenotypes (dbGaP, accession:
   phs003581.v2.p1).
   Chondrocytes were isolated through enzymatic digestion of cartilage
   tissues and maintained in a confluent monolayer culture to retain the
   chondrocyte phenotype. Cultured cells were treated with
   phosphate-buffered saline (PBS) as a resting condition or 1 µM
   recombinant human fibronectin fragment (FN7-10; also known as FN-f) for
   18 h each, simulating OA-like conditions. OA samples underwent the same
   procedures, except for FN-f treatment. DNA extraction was performed
   using the QIAamp DNA mini kit (Qiagen, #51304), and genotyping
   performed using the Infinium Global Diversity Array-8 v.10 Kit
   (Illumina #20031669). Genotype data were processed with comprehensive
   quality control using PLINK (v1.90b3.45)^[183]64, including removal of
   variants with excessive missing data (> 10%), deviation from
   Hardy–Weinberg equilibrium (p < 1 × 10^−6), and low minor allele
   frequency (< 1%). Sample relatedness was assessed through
   identity-by-descent analysis, retaining only unrelated individuals.
   Population structure was evaluated using principal component analysis
   with EIGENSTRAT (v8.0.0) in combination with 1000 Genomes reference
   data^[184]65,[185]66. Following quality control, data were imputed
   against the TOPMed reference panel with Engle2 (v2.4), yielding
   approximately 9.7 million high-quality autosomal variants^[186]67. RNA
   extraction used the RNeasy kit (Qiagen #74104) with on-column DNase
   digestion. RNA integrity was assessed with the Agilent TapeStation
   4150. The New York Genome Center performed RNA-seq library preparation
   and sequencing.
   No sex-based analysis was performed because the study design was not
   explicitly set up to examine sex differences. The disproportionate sex
   distribution between groups would have limited statistical power for
   meaningful comparisons.
RNA-seq processing and quality control
   RNA-seq data libraries averaged approximately 101 million paired-end
   reads (2 × 100 bp) per sample. Low-quality reads and adapters were
   trimmed using TrimGalore! (v0.6.7)^[187]68, followed by FastQC
   (v0.11.9)^[188]69 quality control. Trimmed fastqs were aligned to the
   GENCODE release 45 (GRCh38.p14) reference genome using the STAR aligner
   (v2.7.7a)^[189]70. Gene expression levels were estimated using Salmon
   (v1.9)^[190]71 with –seqBias, –gcBias, and –validateMappings flags,
   assembly of GENCODE version 45 transcript sequences. The
   tximeta^[191]72 R package facilitated gene-level scaled transcript
   analysis. RNA signal tracks for PBS resting, FN-f stimulated
   chondrocytes, and OA donors were created using deepTools
   (v3.5.4)^[192]73 and merged by condition. We evaluated genotype
   consistency between RNA-seq and genotyping array data using VerifyBamID
   (v1.1.3)^[193]74. Two genotyping sample swaps were corrected. Samples
   with FREEMIX and CHIPMIX scores above 0.2 after correction were
   omitted. This process was also applied to OA donor samples. The final
   dataset included 101 PBS (resting) and 101 FN-f treated chondrocyte
   samples from non-OA donors, and 16 untreated OA donor samples.
Splicing quantification of introns usage
   We quantified alternative splicing events using LeafCutter
   (v0.2.9)^[194]75, which allows for the detection of both annotated and
   novel splicing events. Exon-exon junctions were extracted from uniquely
   mapped reads using RegTools (v1.0.0)^[195]76 with default settings. To
   determine how many of our differential splice junctions contained a
   canonical motif, we collated the motif annotations reported by the STAR
   aligner. LeafCutter analysis was performed with a minimum threshold of
   100 supporting reads per intron and a maximum intron length of 500,000
   base pairs, using the GENCODE release 45 (GRCh38.p14) reference genome
   and GTF as reference. LeafCutter defines intron clusters representing
   alternative splicing choices by grouping overlapping introns.
   Differential splicing events were annotated by LeafCutter based on
   overlap with the GENCODE v45 transcriptome as annotated, including
   cryptic 5′, cryptic 3′, cryptic unanchored, and novel annotated pairs.
   We calculated the PSI for each intron based on its relative usage
   within its cluster. Intron excision ratios underwent standardization
   across individuals and quantile normalization across samples. Global
   alternative splicing principal components were calculated using the
   prepare_phenotype_table.py script from LeafCutter, using raw counts of
   intron excision from the intron clustering step.
Differential splicing analysis
   Differential splicing analysis was performed using LeafCutter_ds.R,
   considering only introns used in at least 10% of samples and with at
   least 10 samples per group having sufficient coverage. We compared
   PBS-resting vs. FN-f-stimulated chondrocytes and PBS-resting vs.
   osteoarthritic (OA) chondrocytes. PSI values were normalized using a
   custom method that accounts for cluster-wise proportions and handles
   missing data. The batch correction was determined by correlations
   between technical confounders and the top 10 principal components of
   global splicing in resting and FN-f stimulated chondrocytes. We
   identified that the RNA extraction kit batch and fragment batch were
   the primary sources of batch effects in our data. To address these
   issues, along with donor-specific variations, we applied the
   removeBatchEffect() function from the limma package^[196]18. For the
   final differential splicing analysis, we employed a mixed-effects
   model:
   [MATH:
   model.m<
   mi>atrix~condi<
   /mi>tion+co
   mi>variates+(1∣D
   mi>onorID<
   mo>) :MATH]
   1
   Covariates included RNA shipping date, RNA extraction kit batch, and
   fragment batch. Significantly spliced intron junctions were identified
   using a threshold of adjusted p-value (padj) < 0.05 (Benjamini–Hochberg
   procedure for multiple test correction) and absolute PSI difference
   (|∆PSI|) > 0.15.
Characterization of differentially spliced genes
   Significantly spliced intron junctions (padj <0.05 and |∆PSI| > 0.15)
   were identified in comparisons of PBS-resting vs. FN-f stimulated
   chondrocytes and PBS-resting vs. OA chondrocytes. We performed pathway
   (KEGG, Reactome source) and GO analyses on the associated genes using
   the findMotifs.pl script from the HOMER^[197]77 software suite (v4.11).
   Enriched GO terms (p < 0.01) and KEGG and Reactome pathways (p < 0.01)
   were identified, with GO terms further refined based on semantic
   similarity using rrvgo (v1.14.2)^[198]78. Protein-protein interaction
   networks were constructed using the STRING database (v12.0)^[199]79,
   considering only high-confidence interactions (combined score > 0.9).
Comparative splicing analysis in OA and FN-f chondrocytes
   We compared differentially spliced intron junction-associated genes
   between PBS-resting versus FN-f stimulated and PBS-resting versus OA
   chondrocytes. Both comparisons used the same thresholds (padj < 0.05
   and |∆PSI| > 0.15). To assess the relationship between OA-associated
   and FN-f-induced splicing changes, we examined the PSI in OA for genes
   showing differential splicing in response to FN-f in chondrocytes. The
   significance of these comparisons was evaluated using the Wilcoxon
   test, with p < 0.01 considered statistically significant.
RNA isolation and RT-qPCR analysis of SNRNP70 splicing
   For RT-qPCR validation experiments, a separate set of primary human
   chondrocytes was isolated from non-OA donors following the same
   protocols described in our dataset information section. Cells were
   treated with PBS or FN-f for 18 h under identical conditions to our
   RNA-seq experiments. Total RNA was extracted using the RNeasy kit
   (Qiagen #74104) with gDNA eliminator columns, followed by the Turbo
   DNA-free kit (cat AM1907). RNA was reverse transcribed using the
   High-Capacity RNA-to-cDNA kit (cat #4387406) according to the
   manufacturer’s instructions. qPCR was performed using Power-up SYBR
   Green Master Mix (cat A25742) on a 96-well format Applied Biosystems
   QuantStudio™ 6 Flex machine (Applied Biosystems, Foster City, CA, USA).
   Custom SYBR green primers specific to the SNRNP70 alternative splicing
   junctions were designed. These included primers targeting the junction
   between exon 7 and alternative exon 8 (Forward:
   5′-CGAGACATGCACTCCACCAC-3′, Reverse: 3′-CTGAGGTGTTGTAAGGGGAGC-5′), the
   junction between exon 7 and regular exon 8 (Forward:
   5′-CGAGACATGCACTCCGCTTA-3′, Reverse: 3′-CTCCAGCCCTTCACGGTTC-5′), and a
   constitutive region spanning exons 3–5 (Forward:
   5′-GCGCATGGAGAGGAAAAGAC-3′, Reverse: 3′-CCCCTGAGCATTGGGATCAT-5′).
   Expression levels were normalized to the constitutive exon region
   (exons 3–5). PCR cycling conditions were 95 °C for 2 min; 40 cycles of
   95 °C for 15 s, 60 °C for 60 s. To confirm amplicon specificity and
   expected product sizes, PCR products were visualized on 1.5% agarose
   gels stained with SYBR Safe DNA Gel stain (Thermo Scientific, cat
   [200]S33102), with all products showing a single amplicon at the
   predicted size.
Western blot analysis
   Primary human chondrocytes were treated with PBS or FN-f for 18 h as
   described above. The cells were lysed in Cell Lysis Buffer (Cell
   Signaling Technology, cat 98035), supplemented with Halt Protease and
   Phosphatase Inhibitor Cocktail (ThermoFisher, cat 78440) and
   phenylmethanesulfonyl fluoride solution (Cell Signaling cat 8553S).
   Protein concentration was determined using the Micro BCA Protein Assay
   Kit (Thermo Fisher, cat 23235), according to the manufacturer’s
   instructions. Protein (12 µg per lane) was mixed with 4× Laemmli buffer
   containing 2-mercaptoethanol, heated at 95 °C for 5 min, and separated
   on polyacrylamide gels prepared with the TGX Stain-Free FastCast
   Acrylamide Kit (Bio-Rad, cat 1610180), using 10% ammonium persulfate
   and TEMED. The separated proteins were transferred to 0.2 μm PVDF
   membranes using Trans-blot Turbo transfer buffer (Bio-Rad, cat
   10026938). After blocking in 8% non-fat dry milk in Tris-buffered
   saline with 0.1% Tween-20 (TBST) for 90 min at room temperature, the
   membranes were incubated overnight (18 h) at 4 °C with SNRNP70
   Recombinant Monoclonal Antibody (ThermoFisher, cat MA5-55557, 1:1000
   dilution) and GAPDH (Trevigen, cat 2275-PC-100) as the loading control.
   After washing with TBST, the membranes were incubated with secondary
   antibody (Cell Signaling Technology, cat 7074 and cat 7076, 1:1000
   dilution) for 90 min at room temperature. Protein bands were visualized
   using Radiance Peroxide and ECL substrate (Bio-Rad, cat 1705062). Band
   intensities were quantified using FIJI, and SNRNP70 expression was
   normalized to GAPDH. A Student’s t-test was used to calculate the
   p-value. Full western blots are provided in Supplementary Fig. [201]8.
sgRNA design and RNP complex formation
   Two custom Alt-R crRNAs targeting SNRNP70 (5′-TTGGTTGAGGACAGCCCCCT-3′
   and 5′-TATCACCCCACTGCCCGCTG-3′) were designed to target the plus and
   minus strands, respectively. These crRNAs, synthesized by Integrated
   DNA Technologies (IDT), exclude the PAM sequence. Guide RNA (gRNA)
   duplexes were formed by combining 1.5 µL each of Alt-R tracrRNA (IDT,
   cat 1073190) and crRNA in PCR tubes. The mixtures were annealed at
   95 °C for 5 min and cooled to room temperature. Ribonucleoprotein (RNP)
   complexes were assembled immediately before nucleofection by combining
   1.2 µL of annealed gRNA with 2.6 µL of Alt-R Cas9 Nuclease V3 (IDT, cat
   1081058).
Nucleofection of primary human chondrocytes
   Primary human chondrocytes isolated from macroscopically normal
   cartilage of cadaveric ankles were cultured in DMEM/F-12 medium
   supplemented with 10% FBS and antibiotics. Approximately 3 × 105 cells
   per condition underwent nucleofection using the Lonza Nucleofector
   system (program ER100). Cells were resuspended in 16 µL nucleofector
   solution, combined with 5 µL RNP complex (or PBS for mock control) and
   1 µL Cas9 Electroporation enhancer (IDT, cat 1075916), then
   electroporated. Post-nucleofection, cells were cultured in an
   antibiotic-free medium with 20% FBS at 37 °C and 5% CO[2] for at least
   one week.
Genome editing and colony screening
   Genomic DNA was isolated using a column-based extraction method,
   following the QIAGEN DNeasy Blood & Tissue Kit protocol (QIAGEN, Cat
   69504). PCR screening targeted the SNRNP70 alternative exon 8 region
   using specific primers (Forward: 5′-CTCCTCCCTCTGTTTCTGATG-3′, Reverse:
   5′-CAGGAAAGGGGAGTCGTAGAG-3′). PCR reactions (20 µL total) contained
   1–5 µL genomic DNA, 10 µL Platinum II hot start Green master mix
   (Invitrogen, cat 14001012), and 0.5 µL each of forward and reverse
   primers (10 µM stocks). PCR cycling conditions were 94 °C for 2 min; 40
   cycles of 94 °C for 15 s, 57 °C for 30 s, 68 °C for 45 s; final
   extension at 68 °C for 2 min. Products were analyzed by 1.5% agarose
   gel electrophoresis and stained with SYBR Safe DNA Gel stain (Thermo
   Scientific, cat [202]S33102). PCR final product gel electrophoresis is
   provided in Supplementary Fig. [203]8.
Single-cell colony selection and expansion
   Edited cells were seeded at ~200 cells per 6 cm^2 dish. After 20 days,
   individual colonies were manually isolated. Selected colonies were
   expanded from 24-well plates and plated as passage 1 cells, then
   cultured in media supplemented with 1 ng/mL TGF-β1 (Life Technologies,
   cat PHG9214) and 5 ng/mL bFGF (Life Technologies, cat PHG0264) to
   expand. All 46 selected single colonies were confirmed to have edits by
   PCR screening as described above. Further SNRNP70 alternative exon 8
   deletion was confirmed by Sanger sequencing of PCR products from both
   wild-type and potential heterozygous knockdown (Het-KD) clones. For
   Het-KD clones, the 512 bp band was gel-extracted. PCR products were
   purified using Exonuclease I (EXO1) and Shrimp Alkaline Phosphatase
   (SAP) treatment, following Eurofins Genomics sample preparation
   guidelines. Purified products were sequenced using SimpleSeq tubes
   (Eurofins Genomics: [204]https://eurofinsgenomics.com/). To minimize
   potential clone-specific effects and ensure that observed changes were
   due to heterozygous deletion of SNRNP70 alternative exon 8 rather than
   individual clonal variation, multiple independently derived clones for
   each genotype were pooled. Five wild-type clones were combined into one
   group, and four heterozygous knockout clones into another. Each group
   was plated into six wells: two for sequencing, one as a backup, and two
   for RNA-seq experiments.
RNA sequencing of Het-KD cells
   Cells were expanded to 500,000 cells per well in six-well plates. Total
   RNA was extracted from two sequencing replicates of each cell
   population using the RNeasy kit (Qiagen #74104) with on-column DNase
   digestion. RNA quality and quantity were assessed using the Qubit RNA
   High sensitivity assay (Thermo Fisher Scientific #[205]Q32582) and the
   Agilent TapeStation 4150 system. Library preparation was performed
   using the KAPA RNA HyperPrep Kit with RiboErase (Roche #KK8560)
   following the manufacturer’s protocol. Libraries were quantified using
   the KAPA Library Quantification Kit (Roche #07960298001). Sequencing
   was conducted on a NextSeq 2000 system using NextSeq™ 1000/2000 P1
   Reagents (100 cycles, Illumina #20074933), generating an average of
   24.5 million paired end reads (2 × 50 bp) per sample.
Edited cell quantification and differential expression
   RNA-seq data from edited cells were processed using a pipeline similar
   to the one described for the PBS-resting, FN-f-stimulated, and OA
   chondrocyte RNA-seq data. Briefly, quality control with FastQC
   (v0.11.9)^[206]69 and reads were aligned to the GENCODE release 45
   (GRCh38.p14) reference genome using the STAR aligner (v2.7.7a)^[207]70.
   Gene expression levels were quantified using Salmon (v1.9)^[208]71 with
   –seqBias, –gcBias, and –validateMappings flags, utilizing GENCODE
   version 45 transcript sequences. Gene-level scaled transcript analysis
   was facilitated by the tximeta^[209]72 R package. Differential gene
   expression analysis between wild-type and SNRNP70 alternative exon 8
   heterozygous knockout cells was performed using DESeq2^[210]80. We
   filtered out lowly expressed transcripts, retaining only those with at
   least 10 counts in 10% or more of the samples. Differential expression
   was defined by an FDR-adjusted p-value < 0.01 and absolute log[2] fold
   change > 2, using the apeglm method for fold change shrinkage.
SNRNP70 Exon-level quantification
   Exon-level counts for SNRNP70 were generated using featureCounts
   (subread v2.0.6)^[211]81. The resulting counts were analyzed using the
   EBSEA (v1.33.0)^[212]82 R package, which performs statistical testing
   at the exon level before aggregating results to the gene level. This
   approach increases statistical power while accounting for exon
   dependence using empirical Brown’s method^[213]83. We considered exons
   with p-value < 0.05 as significantly differentially used between
   wild-type and heterozygous knockout samples.
Characterization of edited cells
   Differentially expressed genes (log[2]FC > 2 and FDR-adjusted
   p-value < 0.01) were analyzed for KEGG and Reactome pathway enrichment,
   as well as GO terms. Genes were split based on up or downregulation in
   heterozygous knockout cells compared to wild-type. Enriched pathways
   and GO terms (p < 0.01) were identified using the findMotifs.pl script
   from HOMER (v4.11)^[214]77. GO terms were further refined based on
   semantic similarity using rrvgo (v1.14.2) with the following
   parameters: simThresh = 0.7, simMethod = “Rel”^[215]78.
Comparative analysis with FN-f stimulation
   We compared the differential genes between SNRNP70 alternative exon 8
   heterozygous knockout cells and controls with those observed in
   FN-f-stimulated chondrocytes (from our previous PBS-resting versus
   FN-f-stimulated analysis). Log[2] fold changes were examined for genes
   showing differential expression in both conditions. The significance of
   these comparisons was evaluated using the Wilcoxon test, with p < 0.01
   considered statistically significant.
Splicing quantitative trait loci (sQTL) analysis
   We performed sQTL analysis to identify genetic variants associated with
   alternative splicing events in chondrocytes under resting (PBS) and
   osteoarthritis-like (FN-f stimulated) conditions. This analysis was
   built upon the differential splicing using the same PSI values derived
   from LeafCutter^[216]75. We tested genetic variants within ±100 kb of
   each intron cluster for association with PSI values. This window size
   was chosen based on previous studies that have demonstrated the
   enrichment of sQTL SNPs in proximal regions of splice sites^[217]84.
   The ±100 kb range allows for the capture of both proximal and distal
   regulatory elements that may influence splicing^[218]85. We applied
   stringent filtering criteria for variant inclusion using PLINK
   (v1.90b3) to ensure robust associations^[219]64. For autosomal
   chromosomes, we retained variants with either at least 5 heterozygous
   donors or at least 2 homozygous minor allele donors. To account for
   population structure, we calculated genotype principal components (PCs)
   using bcftools (v1.10) and PLINK (v1.90b3)^[220]64,[221]86. Our models
   included the first 10 principal components (PCs), explaining
   approximately 60% of the genetic variance as covariates. We employed
   QTLtools (v1.3), which extends FastQTL functionality, for sQTL
   mapping^[222]30. We incorporated covariates to control for known
   technical variables and hidden factors, optimizing selection through
   iterative testing of global splicing principal components (PCs) 1–20.
   The final sQTL models were:
   For PBS samples:
   [MATH: PSI~S
   NP+10PCsofgenot
   mi>ype+5PCsofgloba
   mi>lsplic
   mi>ing+Fragm
   entBatch
   mi>+RNAextra
   mi>ctionKitBatch
   mi>+SequencingBatch
   mi> :MATH]
   2
   For FN-f samples:
   [MATH: PSI~S
   NP+10PCsofgenot
   mi>ype+4PCsofgloba
   mi>lsplic
   mi>ing+Fragm
   mi>entBatch
   mi>+RNAextra
   mi>ctionKitBatch
   mi>+SequencingBatch
   mi> :MATH]
   3
   To determine significant associations, we implemented a
   permutation-based approach using QTLtools^[223]30. We performed 1000
   permutations for each phenotype to characterize the null distribution
   of associations empirically. This permutation scheme approximates the
   tail of the null distribution using a beta distribution. We used the
   qtltools_runFDR_cis.R script to implement the Storey & Tibshirani False
   Discovery Rate (FDR) procedure, identifying significant sQTLs at 5%
   FDR^[224]87. The resulting nominal p-value threshold range was 9.99e-6
   to 0.0001 for both PBS and FN-f conditions. To identify multiple
   independent sQTLs per splice intron junction, we performed conditional
   analysis using QTLtools^[225]30. This method uses a stepwise linear
   regression approach to discover additional independent signals while
   accounting for multiple testing. Secondary signals are variants
   identified after conditioning on the primary signal (lead variant),
   tertiary signals after conditioning on both primary and secondary
   signals, and quaternary signals after conditioning on all previous
   signals, with each level representing an additional independent genetic
   effect. The process involves establishing a nominal p-value threshold
   for each phenotype, iteratively identifying the most robust
   associations, and assigning nearby variants to independent signals. For
   genes with multiple sQTLs identified through conditional analysis, we
   further assessed allelic heterogeneity by calculating pairwise LD using
   PLINK (v1.90b3)^[226]64 and classified variants as independent if they
   had R^2 < 0.7. This allowed us to systematically identify genes with
   multiple independent genetic effects on splicing.
sQTLs and eQTLs overlap analysis
   To investigate the relationship between genetic regulation of splicing
   and gene expression, we analyzed the overlap between our sQTLs and our
   previously published eQTLs derived from the same dataset^[227]9. We
   identified variants in high LD (R^2 > 0.7) between lead sQTLs and
   published eQTLs affecting the same gene. For overlapping variants, we
   performed conditional analysis to determine if sQTL effects were
   independent of eQTL effects. We constructed two linear models using the
   lme4 package in R:
   [MATH: Nullhypot
   mi>hesis=PSI~sQTLgenot
   mi>ype+covariates :MATH]
   4
   [MATH:
   Alterna<
   mi>tivehypot
   mi>hesis=PSI~sQTLgenot
   mi>ype+eQTLgenot
   mi>ype+covariates :MATH]
   5
   We used ANOVA to compare these models and applied the FDR correction to
   determine significance at a 5% FDR.
Condition-specific and response sQTL analysis
   To assess condition-specific effects and potential genetic
   interactions, we applied inverse normal transformation of PSI ratios.
   This transformation was necessary to normalize the distribution of PSI
   values and reduce the impact of outliers, ensuring robust statistical
   analysis. We then conducted interaction testing using linear
   mixed-effects models, comparing reduced models (without interaction
   terms) to full models (including interaction terms) for each
   sQTL-splice intron junction pair. To investigate genetic variants
   influencing splicing in a condition-specific manner, we constructed two
   linear mixed models using the lme4 package^[228]31 in R:
   [MATH: Nullhypot
   mi>hesis=PSI~genotype+co
   variate<
   mi>s+condition+(<
   mn>1∣DonorID) :MATH]
   6
   [MATH:
   Alterna<
   mi>tivehypot
   mi>hesis=PSI~genotype+co
   variate<
   mi>s+condition+genotype:
   conditio
   n+(1∣
   DonorID) :MATH]
   7
   Where covariates included the same technical factors and PCs used in
   the standard sQTL mapping, the condition was coded as 0 (PBS) or 1
   (FN-f), and (1|DonorID) accounted for donor-specific random effects. We
   assessed the significance of the genotype:condition term using ANOVA,
   with p < 0.05 considered significant for the interaction effect.
   To identify high-confidence condition-specific sQTLs, we applied
   additional filters: (1) sQTLs detected only in one condition (PBS or
   FN-f), (2) at least 5 donors per minor allele to ensure robust
   estimation, (3) Absolute beta difference ≥ 0.2 between conditions,
   indicating a substantial change in effect size.
RNA binding protein motif analysis
   We selected 140 RBPs from the ENCODE project, focusing on all eCLIP-seq
   data generated for K562, HepG2, and SM-9MVZL cell lines^[229]88. The
   eCLIP-seq data were processed, aligned to the GRCh38 genome assembly,
   and provided in bed format. We used the Functional enrichment of
   molecular QTLs (Fenrich) module from QTLtools to evaluate the overlap
   between sQTLs and RBP binding sites^[230]30,[231]89. A permutation
   scheme was employed to maintain the distribution of functional
   annotations and molecular QTLs around molecular phenotypes. We
   performed 1000 permutations to estimate the probability of the observed
   overlaps of the lead sQTLs relative to expectation. For sQTLs
   overlapping with RBP binding sites, we calculated the Pearson
   correlation between the RBP gene expression (from the same donor
   RNA-seq data) and the spliced intron junction usage pattern across all
   samples. We considered correlations with R^2 > 0.15 and p-value < 0.05
   as significant. We performed KEGG and Reactome pathway enrichment
   analysis on the genes associated with sQTLs that overlap RBP binding
   sites and show a significant correlation with RBP expression. Using the
   findMotifs.pl script from the HOMER software suite (v4.11), the KEGG
   and Reactome pathways (p < 0.01) were identified.
Colocalization between sQTL and OA GWAS
   We conducted colocalization analysis between our sQTL results and OA
   GWAS signals. Summary statistics for 11 OA phenotypes were obtained
   from Boer et al., encompassing 826,690 individuals of European and East
   Asian descent (Musculoskeletal Knowledge Portal,
   [232]https://msk.hugeamp.org/). These statistics were converted to hg38
   coordinates for consistency with our analysis. Overlapping signals were
   identified by selecting lead sQTL variants in moderate LD (r^2 > 0.5)
   with lead GWAS variants. LD was calculated using PLINK (v1.90b3.45)
   with parameters –ld-window 200000 –ld-window-kb 1000, based on the 1000
   Genomes European reference panel phase 3, chosen due to the East Asian
   population sample size > 4000 individual^[233]4,[234]64,[235]65. We
   defined a window encompassing the sQTL and GWAS lead variants for each
   pair of overlapping signals, extending by 250 kb in each direction. To
   ensure consistent identification, we mapped variants to rsIDs using
   dbSNP 155, querying dbSNP VCF files with bcftools and custom scripts to
   match variants based on chromosome, position, and alleles.
   Colocalization analysis was performed using the Bayesian method coloc
   package, implemented in R^[236]90. The coloc.abf function was used to
   calculate posterior probabilities for various hypotheses, including H3
   (association with sQTL only) and H4 (association with both sQTL and
   GWAS traits). Signals with a posterior probability for H4 (PP H4) > 0.7
   were considered as evidence of significant colocalization. We also
   calculated the ratio (PP H4)/(PP H3 + PP H4) to quantify the strength
   of colocalization evidence. This analysis was conducted separately for
   sQTLs identified in PBS and FN-f conditions to examine
   condition-specific effects on OA risk.
Visualization
   Protein-protein interaction networks were visualized using Cytoscape
   (v3.9)^[237]91. Heatmaps were generated using the ComplexHeatmap R
   package (v3.19)^[238]92. Genomic signal tracks and locus zoom plots
   were created with plotgardener (v1.9.4). All other plots were produced
   using ggplot2 in R^[239]93. These visualizations were generated using R
   version 4.3.
Reporting summary
   Further information on research design is available in the [240]Nature
   Portfolio Reporting Summary linked to this article.
Supplementary information
   [241]Supplementary Information^ (2.4MB, pdf)
   [242]41467_2025_63299_MOESM2_ESM.docx^ (20.5KB, docx)
   Description of Additional Supplementary Files
   [243]Supplementary Data 1^ (87.4KB, pdf)
   [244]Supplementary Data 2^ (183.1KB, xlsx)
   [245]Supplementary Data 3^ (52.4KB, xlsx)
   [246]Supplementary Data 4^ (83.6KB, xlsx)
   [247]Supplementary Data 5^ (2.3MB, xlsx)
   [248]Supplementary Data 6^ (60.2KB, xlsx)
   [249]Supplementary Data 7^ (19.6KB, xlsx)
   [250]Supplementary Data 8^ (13.4KB, xlsx)
   [251]Reporting Summary^ (1.6MB, pdf)
   [252]Transparent Peer Review file^ (898.3KB, pdf)
Acknowledgements