Abstract
   Hutchinson–Gilford progeria syndrome, caused by a mutation in the LMNA
   gene, leads to increased levels of truncated prelamin A, progerin, in
   the nuclear membrane. The accumulation of progerin results in defective
   nuclear morphology and is associated with altered expression of linker
   of the nucleoskeleton and cytoskeleton complex proteins, which are
   critical for nuclear signal transduction via molecular coupling between
   the extranuclear cytoskeleton and lamin‐associated nuclear envelope.
   However, the molecular mechanisms underlying progerin
   accumulation‐induced nuclear deformation and its effects on
   intranuclear chromosomal organization remain unclear. Here, the
   spatiotemporal evolution of nuclear wrinkles is analyzed in response to
   variations in substrate stiffness using a doxycycline‐inducible
   progerin expression system. It is found that cytoskeletal tension
   regulates the onset of progerin‐induced nuclear envelope wrinkling and
   that the molecular interaction between SUN1 and LMNA controls the
   actomyosin‐dependent attenuation of nuclear tension. Genome‐wide
   analysis of chromatin accessibility and gene expression further
   suggests that an imbalance in force between the intra‐ and extranuclear
   spaces induces nuclear deformation, which specifically regulates
   progeria‐associated gene expression via modification of
   mechanosensitive signaling pathways. The findings highlight the crucial
   role of nuclear lamin–cytoskeletal connectivity in bridging nuclear
   mechanotransduction and the biological aging process.
   Keywords: actomyosin contractility, chromatin remodeling,
   heterochromatin, Hutchinson–Gilford progeria syndrome, LINC complex,
   mechanosensation, nuclear deformation, nuclear tension, nuclear
   wrinkling, progerin, SUN1
     __________________________________________________________________
   The premature aging‐related progerin leads to defective nuclear
   morphology and is associated with disrupted molecular coupling between
   the extranuclear cytoskeleton and lamin‐associated nuclear envelope. It
   is discovered that progerin expression reduces nuclear tension, forms
   nuclear wrinkling, and enhances chromatin dynamics, thereby regulating
   progerin‐induced mechanosensitive signaling pathways.
   graphic file with name ADVS-12-2502375-g003.jpg
1. Introduction
   Hutchinson–Gilford progeria syndrome (HGPS), a premature aging disease,
   is caused by a de novo point mutation (G608G; GGC > GGT) in the LMNA
   gene encoding lamins A and C, creating a truncated prelamin A form
   lacking 50 amino acid residues near the C terminus, commonly referred
   to as Δ50 lamin A or progerin.^[ [42]^1 ^] These amino acids include a
   cleavage site for the zinc metallopeptidase STE24 (ZMPSTE24), which is
   crucial in the formation of mature lamin A by removing the farnesyl
   group; loss of this cleavage site leads to a persistent farnesylation
   of progerin.^[ [43]^2 ^] Accumulation of progerin in the nuclear
   envelope (NE) compromises the structural and functional integrity of
   the nucleus, resulting in an abnormal nuclear shape, a hallmark of
   HGPS.^[ [44]^3 ^] For instance, fibroblasts obtained from patients with
   HGPS display defective nuclear shapes, including invaginations,
   lobulations, and wrinkles.^[ [45]^3 , [46]^4 ^] Furthermore,
   overexpression of Δ50 lamin A in primary dermal fibroblasts induces
   abnormalities in nuclear morphology similar to those observed in HGPS
   cells.^[ [47]^5 ^]
   Progerin expression also leads to altered molecular connections between
   the nucleus and cytoskeleton,^[ [48]^6 ^] further resulting in enhanced
   sensitivity to mechanical stress^[ [49]^7 ^] and reduced force
   propagation from the extracellular matrix (ECM) into the cell.^[ [50]^8
   ^] ECM stiffness, which reflects tissue‐specific biomechanical
   properties, is detected by transmembrane receptor proteins, such as
   integrins,^[ [51]^9 ^] which mediates the formation of focal adhesion
   complexes. These complexes transmit mechanical signals into
   intracellular biochemical pathways, thereby inducing diverse cellular
   responses.^[ [52]^10 ^] Moreover, these adhesion complexes promote the
   polymerization of intracellular actin cytoskeletal networks, where
   actomyosin contractility‐induced mechanical forces allow cells to sense
   and respond to the biomechanical characteristics of their surrounding
   environment.^[ [53]^11 ^] These force‐transmitting molecular
   connections are largely mediated by the linker of the nucleoskeleton
   and cytoskeleton (LINC) complexes.^[ [54]^12 ^] The LINC complex is
   composed of a SUN (Sad1, UNC‐84) domain located in the inner nuclear
   membrane and a KASH (Klarsicht, ANC‐1, and Syne Homology) domain
   located in the outer nuclear membrane.^[ [55]^13 ^] Nuclear envelope
   spectrin‐repeat proteins (nesprins), consisting of actin‐binding
   N‐terminal and KASH domain‐binding C‐terminal domains, also bind to SUN
   proteins in the perinuclear space between the nuclear membranes.^[
   [56]^14 ^] Thus, the LINC complex transmits biophysical stimuli into
   the nucleus through nesprin‐mediated molecular connections between the
   actin cytoskeleton and SUN proteins.^[ [57]^15 ^] In particular,
   LINC‐mediated cytoskeletal tension plays a key role in cellular
   mechanoresponses,^[ [58]^16 ^] where actomyosin contractility and
   force‐dependent reorganization of the cytoskeletal architecture remodel
   LINC‐associated protein components to transmit extracellular physical
   stimuli into the nucleus.^[ [59]^17 ^]
   Lamins interact with heterochromatic genomic regions via
   lamina‐associated domains (LADs) that are enriched with repressive
   histone marks, such as trimethylated histone H3 at lysine 9 (H3K9me2/3)
   and H3K27me3.^[ [60]^18 ^] Transcriptional repression of genes within
   such heterochromatic regions is attributed to the loss of DNA–nuclear
   lamina interactions that contribute to chromatin condensation and the
   expression of H3K9me2/3.^[ [61]^19 ^] Thus, lamin‐associated
   intranuclear chromatin organization correlates with nuclear stiffening,
   which is crucial for maintaining nuclear structural integrity in
   response to mechanical stress.^[ [62]^20 ^] Because NE‐associated SUN
   proteins directly bind chromatin to the nuclear periphery,^[ [63]^21 ^]
   impaired LINC complexes hinder intracellular force transmission, which,
   in turn, disrupts perinuclear cytoskeleton‐dependent mechanosensing of
   the extracellular microenvironment. Ultimately, mechanosensitive
   nuclear deformation compromises intranuclear elasticity distribution,^[
   [64]^22 ^] indicating the remodeling of heterochromatin accessibility.
   Reduced chromatin mobility by the inhibition of actomyosin
   contractility or nuclear detachment from the cytoskeleton further
   highlights the critical role of an intact actomyosin apparatus and the
   LINC complex in mechanical signal transduction to chromatin.^[ [65]^22
   , [66]^23 ^]
   Progerin expression disrupts the LINC complex by altering the
   expression of SUN proteins without affecting the localization of SUN1
   and SUN2 to the nuclear membrane,^[ [67]^24 ^] resulting in altered
   cellular mechanotransduction.^[ [68]^25 ^] In NIH 3T3 fibroblasts
   expressing myc‐tagged progerin and in HGPS fibroblasts, enhanced SUN1
   expression in response to progerin accumulation alters nuclear coupling
   to actin filaments (F‐actin) and microtubules (MT) through nesprin 2,
   resulting in defective nuclear movement and cell polarity.^[ [69]^26 ^]
   SUN1 overexpression in HGPS fibroblasts can further lead to increased
   SUN1–nesprin 2 coupling with MT, thereby inhibiting nuclear movement
   due to its complementary coupling with the actin cytoskeleton.^[
   [70]^26b ^] Consequently, nesprin 2‐mediated actomyosin tension applied
   to the NE is reduced,^[ [71]^27 ^] confirming a decrease in nuclear
   forces in response to SUN1 overexpression in progerin‐expressing cells.
   Moreover, SUN1 silencing in HGPS fibroblasts rescues the deformed
   nuclear morphology,^[ [72]^28 ^] whereas SUN2 depletion fails to
   restore progerin‐induced nuclear deformation.^[ [73]^4 ^] These
   observations highlight the role of SUN1 upregulation in nuclear
   deformation as a molecular hallmark of the pathological signature of
   HGPS.
   Progerin expression increases F‐actin polymerization^[ [74]^29 ^] and
   RhoA activation,^[ [75]^30 ^] resulting in cytoskeletal stiffening.^[
   [76]^7 ^] For instance, F‐actin polymerization is increased in
   mesenchymal stromal/stem cells from the ZMPSTE24 ^−/− HGPS mouse model,
   Z24^−/− mesenchymal stromal cells (MSCs), which exhibit a higher
   elastic modulus than wild‐type MSCs. Increased RhoA activity in Z24^−/−
   MSCs is suppressed by pharmaceutical inhibition of RhoA signaling,
   resulting in reduced nuclear deformation.^[ [77]^30 ^] Consistent with
   the results from Z24^−/− MSCs, increased F‐actin polymerization and
   nuclear deformation in HGPS fibroblasts are reversed by treatment with
   a farnesyltransferase inhibitor.^[ [78]^30 , [79]^31 ^] Elevated
   cytoskeletal tension in progeria cells arises from mechanical stress
   due to both ECM stiffness^[ [80]^32 ^] and increased nuclear rigidity,
   as sustained RhoA signaling promotes F‐actin cytoskeletal stiffness.^[
   [81]^30 ^] Progerin‐induced alterations in nucleocytoskeletal
   connections result in reduced propagation of cytoskeletal forces to the
   nucleus,^[ [82]^8 ^] as determined by particle tracking analysis in
   exogenous progerin‐expressing HeLa cells, human umbilical vein
   endothelial cells, and HGPS fibroblasts.^[ [83]^8 ^]
   Accumulating evidence suggests that nucleus‐responsive forces can alter
   epigenetic modifications through chromatin remodeling;^[ [84]^20 ^]
   mechanosensitive changes in forces applied to the nucleus can alter
   intranuclear heterochromatin reorganization and chromatin
   accessibility. For instance, dermal fibroblasts derived from patients
   with HGPS show changes in DNA methylation, and chromatin accessibility
   is enriched in NE‐associated regions, contributing to the abnormal gene
   expression patterns observed in HGPS.^[ [85]^33 ^] Therefore, progerin
   accumulation leads to widespread alterations in the repressive histone
   mark H3K27me3, disrupting the association between heterochromatin and
   the nuclear lamina in HGPS fibroblasts, resulting in the loss of
   chromosomal compartmentalization.^[ [86]^34 ^]
   Previously, studies on HGPS models have consistently demonstrated
   increased cytoskeletal tension and RhoA activation,^[ [87]^30 ^] which
   are essential to regulate force transmission to the nucleus. In turn,
   nuclear forces alter epigenetic modifications through chromatin
   remodeling.^[ [88]^20 , [89]^33 ^] However, the causal relationship
   between progerin‐induced nuclear–cytoskeletal force transmission and
   nuclear deformation in progerin‐expressing cells, as well as the
   consequent alterations in gene expression remains to be identified.
   While previous studies have shown that nuclear deformation can be
   regulated by overexpressed SUN1, the precise mechanism by which SUN1
   overexpression induces nuclear deformation remains poorly understood.
   Therefore, here, we investigated the biophysical mechanisms that
   regulate nuclear morphological changes in response to progerin
   expression and their downstream effects on gene expression. To this
   end, we developed a human HGPS cell model, where the progerin
   expression was precisely controlled using the doxycycline‐induced
   Tet‐On system, overcoming the constraints of using the primary cells
   from patients with HGPS owing to their limited availability. We
   elucidated the mechanism by which progerin expression alters nuclear
   morphology over time. We further investigated how progerin accumulation
   at the nuclear membrane in HGPS cells affects LINC complex expression
   and the physical connectivity between the actin cytoskeleton and
   nuclear lamina. Furthermore, we examined the impact of these nuclear
   morphological changes on chromatin organization and the aberrant gene
   expression profiles in HGPS. Our study improves our understanding of
   the progerin‐induced alteration of nuclear tension that drives nuclear
   deformation.
2. Results
2.1. Mechanosensing of Substrate Stiffness Modulates Progerin‐Induced Nuclear
Deformation
   HGPS, characterized by abnormal nuclear morphology due to the
   expression of truncated prelamin A (Δ50 LMNA/progerin),^[ [90]^35 ^]
   displays impaired nuclear mechanotransduction.^[ [91]^26b ^] To
   elucidate the molecular mechanism by which progerin accumulation alters
   mechanical signal transduction through nuclear deformation, we
   developed an HGPS model in HeLa cells, where the onset of progerin
   expression was precisely controlled by transfection with the XLone
   plasmid that combined the piggyBac transposon and Tet‐On 3G
   doxycycline‐inducible gene expression system (see the Experimental
   Section and Figure [92]S1, Supporting Information).^[ [93]^36 ^]
   The treatment of Tet‐On gene‐expressing HeLa cells with 2 µg mL^−1
   doxycycline specifically induced progerin expression and dramatically
   increased the population of abnormal nuclei with blebs, lobulation, and
   wavy surface texture, while the overall cell and nuclear size remained
   unchanged (Figure [94]1A–E; Figure [95]S2A, Supporting Information). We
   noted that traditional nuclear morphometry measurements, including
   nuclear area, circularity, and aspect ratio (defined as nuclear
   spreading area, 4π(area)/(perimeter^2), and the ratio of the longest
   axis to the perpendicular shortest axis, respectively), were
   insufficient to quantitatively assess the nuclear deformation induced
   by progerin expression (Figure [96]1B,C; Figure [97]S2B,C, Supporting
   Information). Hence, we assessed the NE wrinkling area by measuring the
   fraction of the nuclear surface area occupied by NE wrinkles to more
   precisely represent progerin‐induced NE‐defective nuclei (Figure
   [98]S2D, Supporting Information). To further confirm that
   doxycycline‐induced progerin expression controls the progression of
   nuclear deformation, we systematically quantified progerin expression
   and NE wrinkling at different time points, where we noted that the
   time‐dependent elevation of progerin expression preceded the increase
   in NE wrinkling (Figure [99]S3, Supporting Information). These results
   demonstrate that our doxycycline‐controlled progerin‐expressing Tet‐On
   HeLa cell line not only replicates the deformed nuclear morphology
   detected in HGPS cells but also enables time‐dependent monitoring of
   progerin expression and nuclear deformation.
Figure 1.
   Figure 1
   [100]Open in a new tab
   Substrate stiffness‐dependent differential evolution of nuclear
   deformation in doxycycline‐inducible progerin‐expressing HeLa cells.
   A–E) Morphological alterations of doxycycline‐controlled Tet‐On HeLa
   cells expressing mutant lamin A protein (Δ50 LMNA/progerin).
   Representative confocal images depict immunofluorescence staining for
   progerin (red), F‐actin (green), and nuclei (DAPI, blue) in
   doxycycline‐untreated control cells (−Dox) or doxycycline‐treated (2 µg
   mL^−1) progerin‐expressing Tet‐On HeLa cells (+Dox). Hemispherical and
   cross‐sectional views of 3D‐rendered nuclei show progerin
   expression‐induced formation of abnormal nuclear morphology. Empty and
   full arrowheads indicate the absence and presence of progerin
   expression, respectively (A). Immunofluorescence intensity‐based
   quantifications of cell area (B), nuclear area (C), progerin expression
   (D), and the fractional occurrence of abnormal nuclear shapes (E) were
   performed in the absence and presence of doxycycline treatment. In
   panel B, 810 and 505 nuclei; in panel C, 138 and 153 nuclei; in panel
   D, 93 and 72 nuclei were analyzed under −Dox and +Dox conditions,
   respectively. For panel E, 70 to 105 nuclei were analyzed, which was
   independently repeated three times per each condition. Error bars
   indicate the standard error of the mean (S.E.M.); an unpaired t‐test
   was applied (***: p < 0.001, NS: not significant). F–G) Substrate
   stiffness‐dependent differential expression of progerin. Progerin
   (red), lamin B1 (green), and nuclei (DAPI, blue) of progerin‐expressing
   Tet‐On HeLa cells were plated on control glass substrates or
   polyacrylamide hydrogel (PAG) substrates with elastic moduli of 34 kPa
   (stiff) and 1.37 kPa (soft) (F). Control glass and stiff PAG substrates
   maintained doxycycline‐inducible progerin expression, while soft PAG
   substrates significantly reduced progerin expression (G). In panel G,
   >50 cells were tested per condition. Error bars indicate the S.E.M.;
   Student's t‐test was applied (***: p < 0.001, NS: not significant).
   H–J) Time‐lapse monitoring of substrate stiffness‐dependent progerin
   expression and nuclear deformation. Progerin intensity and nuclear
   envelope (NE) wrinkling were monitored every 20 min for 36 h in
   mCherry‐progerin‐expressing Tet‐On HeLa cells plated on control glass
   (H), stiff PAG (I), and soft PAG (J) substrates. Yellow dotted lines
   mark the nuclear boundary determined by differential interference
   contrast (DIC) imaging, showing nuclear spreading area. Full and empty
   arrowheads indicate the presence and absence of progerin (red) or NE
   wrinkling (blue), respectively; transparency of the full arrowheads
   represents the magnitude of progerin expression and NE wrinkling;
   oversaturated fluorescence intensity (white) indicates nuclear surface
   wrinkles (H–J). K–P) Quantifying differential onset of progerin
   expression and nuclear deformation in response to changes in substrate
   stiffness. Progerin expression and NE wrinkling were quantified by
   measuring fluorescence intensity (red curves) and the fraction of the
   nuclear spreading area occupied by NE wrinkling area (blue curves)
   after doxycycline treatment (K–M). All values were normalized using the
   formula (x – x [min])/(x [max] – x [min]) to range from 0 (min) to 1
   (max). Progerin expression and NE wrinkling followed extended sigmoidal
   curves, with inflection points for progerin expression at 20 h and NE
   wrinkling at 25 h on control glass and stiff PAG substrates (K,L,N,O),
   which were delayed to 22 and 32 h, respectively, on soft PAG substrates
   (M–O). The time interval between inflection points of progerin
   expression and NE wrinkling extended from 5 h on control glass and
   stiff PAG substrates to 10 h on soft PAG substrates (P). In panels K–P,
   >20 cells were analyzed per condition. Error bars indicate the S.E.M.;
   one‐way analysis of variance (ANOVA) with Tukey's test was used for
   comparisons (***: p < 0.001, NS: not significant).
   Mechanosensing of matrix rigidity alters protein expression by
   remodeling the nucleus–cytoskeletal connection, mediating the
   intracellular mechanical balance between cytoskeletal force and nuclear
   surface tension.^[ [101]^37 ^] Accordingly, we investigated whether
   cells adapted to differential substrate stiffness could modulate
   progerin expression in doxycycline‐activated HeLa cells
   (Figure [102]1F,G). Cells placed on control glass and polyacrylamide
   hydrogel (PAG) substrates, mimicking the in vivo elastic moduli of
   rigid and compliant organs,^[ [103]^38 ^] displayed deformed nuclear
   shapes in response to doxycycline‐induced progerin expression,
   regardless of substrate stiffness (Figure [104]1F). In contrast to
   doxycycline‐untreated cells not expressing progerin,
   doxycycline‐treated cells placed on rigid PAG substrates (E ≈ 34 kPa)
   maintained elevated progerin expression similar to cells placed on
   control glass substrates, while those grown on soft PAG substrates (E ≈
   1.37 kPa) showed a significant reduction in progerin expression
   (Figure [105]1G). This result strongly suggests that substrate
   stiffness‐dependent changes in intracellular tension could regulate
   progerin expression.
   Since doxycycline‐controlled progerin expression led to nuclear
   deformation in a time‐dependent manner (Figure [106]S3, Supporting
   Information), and progerin expression levels were altered by substrate
   stiffness (Figure [107]1F,G), we assessed whether differential
   substrate stiffness could modulate the temporal relationship between
   doxycycline‐induced progerin expression and nuclear deformation. To
   this end, we conducted real‐time live cell monitoring of mCherry‐tagged
   progerin‐expressing Tet‐On HeLa cells placed on substrates with varying
   stiffness (Figure [108]1H–J; Movies [109]S1 and [110]S2, Supporting
   Information). Time‐lapse imaging revealed that doxycycline‐induced
   progerin expression exhibited sigmoidal changes, followed by NE
   wrinkling, which was most evident on control glass substrates, but
   attenuated on reduced substrate stiffness (Figure [111]1H–M).
   Assessment of the inflection time points in the curves further revealed
   that progerin expression and nuclear deformation were delayed in cells
   placed on soft PAG substrates compared to those placed on control glass
   or stiff PAG substrates (Figure [112]1K–M). Specifically, the
   inflection points for progerin expression and NE wrinkling occurred at
   20 and 25 h on control glass and stiff PAG substrates, respectively,
   whereas they were observed at 22 and 32 h on soft PAG substrates
   (Figure [113]1K–M). Consequently, the time interval between the
   inflection points of the curves for progerin expression and NE
   wrinkling was extended from 5 h on control glass and stiff PAG
   substrates to 10 h on soft PAG substrates (Figure [114]1N–P).
   These results indicate that doxycycline‐induced progerin expression,
   followed by NE wrinkling, is highly mechanosensitive to changes in
   substrate stiffness, which further suggests that progerin‐induced
   nuclear deformation can be regulated by intracellular cytoskeletal
   tension applied to the nucleus.
2.2. Mechanosensing of Substrate Stiffness Induces the Differential
Attenuation of Nuclear Tension in Progerin‐Expressing Cells
   Previously, we demonstrated that substrate stiffness‐dependent spatial
   reorganization of lamin A/C is regulated by cytoskeletal tension,^[
   [115]^16 ^] and that 3D morphological alterations of the nucleus in
   response to mechanical stimuli are regulated by nucleus–cytoskeletal
   connections.^[ [116]^39 ^] Because the LINC‐mediated molecular
   connection between the cytoskeleton and nuclear lamina facilitates
   force transmission across the nuclear membrane,^[ [117]^21 ^] we
   hypothesized that the nuclear tension between these elements could
   determine the substrate stiffness‐modulated differential onset of
   progerin‐induced nuclear deformation.
   To directly measure the force applied to the nuclear–cytoskeletal
   connection, we used a fluorescence energy transfer (FRET)‐based NE
   tension sensor module tagging the inner nuclear membrane SUN‐binding
   domain and cytoplasmic F‐actin‐binding domain, mimicking the LINC
   complex associating nesprin 2^[ [118]^27 ^] (Figure [119]S4A,
   Supporting Information). As an elevated FRET ratio in HGPS cells
   implies diminished nuclear tension due to reduced actomyosin
   contractility,^[ [120]^27 ^] we tested whether doxycycline‐induced
   progerin expression could impinge on the substrate stiffness‐dependent
   differential decline of nuclear tension. To monitor time‐dependent
   gradual changes in NE tension, we transiently transfected the nesprin
   tension sensor into progerin‐expressing Tet‐On HeLa cells before
   placing them on substrates of varying stiffness (Figure [121]S4B,C,
   Supporting Information). We confirmed that the transfected nesprin
   tension sensors were specifically localized along the nuclear membrane
   and that their fluorescence intensity was maintained without
   significant decay during time‐lapse imaging of the nuclei every 6 h for
   36 h (Figure [122]S4D,E, Supporting Information).
   To analyze the substrate stiffness‐dependent differential application
   of NE tension in response to doxycycline‐induced progerin expression,
   we masked FRET signals located along the NE (Figure [123]2A–C), thereby
   preventing interference with FRET signals stemming from other regions
   of the cell (see the Experimental Section for details). FRET efficiency
   was calculated by dividing the fluorescence intensity of the acceptor
   by that of the donor after background subtraction within the masked NE
   region, which was differentially color‐coded, where approaching purple
   indicated a decreased FRET ratio due to enhanced NE tension, and
   conversely, approaching red indicated an increased FRET ratio due to
   reduced NE tension (Figure [124]2A–C).
Figure 2.
   Figure 2
   [125]Open in a new tab
   Substrate stiffness‐dependent differential NE tension. A–C)
   Visualization of substrate stiffness‐dependent changes in NE tension
   using the nesprin tension sensor. Tet‐On HeLa cells expressing
   fluorescence‐marker‐untagged progerin were transiently transfected with
   the nesprin tension sensor and plated on control glass or stiff (E ≈ 34
   kPa) or soft (E ≈ 1.37 kPa) PAG substrates. NE tension was analyzed
   every 6 h for 36 h after doxycycline treatment. Binary layers outlining
   the nesprin tension sensor‐localized nuclear membrane were determined
   by creating polygonal hollow masks to exclude fluorescence intensity
   outside the nucleus (top rows). FRET signals were differentially
   color‐coded (bottom rows). Purple and red indicates high FRET
   efficiency. FRET efficiency, representing the inverse of NE tension,
   gradually increased in response to doxycycline‐induced progerin
   expression, but this rate was reduced in cells on soft PAG substrates
   compared to those on control glass and stiff PAG substrates (A and B vs
   C). D–G) Quantification of substrate stiffness‐modulated differential
   NE tension. Column scatter plots represent the time‐dependent increase
   in the FRET ratio in response to doxycycline‐induced progerin
   expression on control glass and stiff or soft PAG substrates. The
   nesprin tension sensor‐based FRET ratio increased significantly after
   24 h in cells on glass (D) and stiff PAG substrates (E), but after 30 h
   in cells on soft PAG substrates (F). Accordingly, the FRET signal in
   cells on glass and stiff PAG substrates was significantly higher than
   in those on soft PAG substrates after 24 h (G). In panels D–G, red bars
   represent the mean ± S.D., and one‐way ANOVA with Tukey's test was
   applied for group comparisons (***: p < 0.001, **: p < 0.05, NS: not
   significant).
   The FRET ratio, representing the opposite of NE tension, significantly
   increased in response to progerin expression, regardless of substrate
   stiffness, after 24 h of doxycycline treatment on control glass and
   stiff PAG substrates (Figure [126]2D,E), while the change was detected
   after 30 h on soft PAG substrates (Figure [127]2F). Interestingly, the
   temporal reduction in NE tension showed no difference between cells
   placed on glass and stiff PAG substrates (Figure [128]2D,E,G).
   Accordingly, a significant difference between the FRET ratios of cells
   placed on rigid and soft substrates was observed after 24 h
   (Figure [129]2G), which is consistent with a previous report showing
   that HGPS cells displayed lower nuclear tension than normal control
   cells.^[ [130]^27 ^]
   Together with our results quantifying the substrate stiffness‐dependent
   differential onset of progerin expression and nuclear deformation
   (Figure [131]1K–P), these data further reveal the temporal relationship
   between the progerin‐induced diminution of NE tension and nuclear
   deformation. We found that the time point at which nuclear tension was
   significantly reduced (i.e., a significantly increased FRET ratio)
   appeared between two distinct time points at which an abrupt increase
   in progerin expression and NE wrinkling were detected. This finding
   strongly supports the hypothesis that progerin induces nuclear
   deformation by reducing nuclear tension. In addition, we noted that the
   time interval between the increment of progerin expression and the
   reduction of nuclear tension, as well as the time interval between the
   reduction of nuclear tension and the increase in NE wrinkling, were
   doubled by relocating cells from the control glass and stiff PAG
   substrates to the soft PAG substrates, i.e., from 4 to 8 h and from 1
   to 2 h, respectively.
   These data indicated that the time required to decrease nuclear tension
   in response to progerin expression was approximately four times longer
   than that required to induce nuclear deformation due to reduced nuclear
   tension. Further, these findings strongly suggest that cytoskeletal
   tension, modulated by changes in substrate stiffness, mediates nuclear
   tension‐dependent nuclear deformation in progerin‐expressing cells.
2.3. Progerin‐Induced Reduction in Nuclear Tension Determines
Mechanosensitive Nuclear Wrinkling
   The temporal onset of progerin expression, reduction in NE tension, and
   nuclear wrinkling in response to doxycycline treatment were delayed by
   relocating cells from stiff to soft substrates (Figures [132]1
   and [133]2), which suggest that not only gene expression driven by
   biochemical stimuli but also its biophysical consequences, such as
   alterations in NE tension and nuclear surface remodeling, are sensitive
   to the mechanical balance between the nucleus and the physical
   environment of the cell. Specifically, we employed a computational
   model to determine whether substrate stiffness‐dependent nuclear force
   mediates the differential evolution of progerin‐induced nuclear
   wrinkling.
   To simulate NE wrinkling, a representative nuclear surface deformation
   induced by progerin expression, we developed a computational model by
   adapting nuclear geometry obtained from 3D‐reconstructed confocal
   images to establish a soft, spherical, elastic, thin shell as the
   initial model structure (Figure [134]3A). The shell was discretized
   into triangular elements using a series of vertices to characterize the
   deformation of the nuclear surface. The potential energy generated by
   the movement of the vertices allows the shell to resist external forces
   (see the Experimental Section for details). Because the FRET analysis
   indicated reduced NE tension (tensile force) at the onset of progerin
   expression (Figure [135]2), we simplified the force transmitting from
   the external environment onto the nucleus, i.e., the cytoskeletal
   force, as a net pressure force. Furthermore, by combining our previous
   report demonstrating that the actin cytoskeleton exerts elevated
   pressure on the nucleus on a stiff substrate,^[ [136]^40 ^] we applied
   varying external pressures to model the nuclear response to changes in
   substrate stiffness, where pressure increases with increasing substrate
   stiffness.
Figure 3.
   Figure 3
   [137]Open in a new tab
   Mechanical model of nucleus wrinkling. A) Construction of a mechanical
   model mimicking progerin‐induced deformation of the nuclear surface.
   Experimental observations of Tet‐On‐inducible progerin expression are
   depicted by 3D reconstructions of confocal images (top) and the
   corresponding mechanical model (bottom), showing the smooth spherical
   shape of progerin‐absent control nuclei (left), a buckyball‐like
   surface configuration at the onset of progerin expression (middle), and
   a folded surface texture due to the progression of nuclear wrinkling
   (right). The color code represents displacement of the nuclear surface.
   B–E) Simulation of time‐dependent nucleus wrinkling in response to
   changes in substrate stiffness. Reduced pressure on the nucleus on
   compliant substrates delays surface wrinkling. The second‐order
   buckling (i.e., from a buckyball‐like pattern to a folded pattern)
   occurs at characteristic time scales of 45, 90, and 105 on stiff,
   medium, and soft substrates, respectively (B). (Inset) The local force
   balance in a membrane microelement, where membrane tension is mimicked
   by an equivalent internal pressure due to surface curvature and
   mechanical equilibrium. Time‐dependent volume changes of the nucleus on
   different stiffness substrates indicate that nuclei on soft substrates
   take the longest time, while those on stiff substrates take the
   shortest, based on the characteristic time to reach a specific volume
   change (C), where ∆V and V [0] represent the volume change and the
   initial volume, respectively, and black circles mark the accelerated
   collapses along the nucleus triggered by the second‐order buckling. The
   nuclear surface tension, calculated from nuclear volume change, shows
   that nuclei on stiff substrates have the highest tension value and
   largest change rate, while nuclei on soft substrates have the lowest
   tension value and smallest change rate (D), where the stretching force
   F [e] was applied to characterize the membrane tension F [memb], F [0]
   represents the unit characteristic force, and the black cross indicates
   the breakdown of computational model due to the contact of the membrane
   under large deformation, respectively. Increasing internal pressure
   (i.e., enhanced membrane tension), corresponding to a greater 𝐾[𝑉]
   value, indicating a stronger resistance to external pressure reduces
   nuclear volume change (E), indicating that reduced nuclear tension on
   soft substrates delays nuclear wrinkling compared to nuclei on stiff
   substrates.
   The nuclear mechanical model revealed that shell wrinkling progresses
   in two stages: the smooth surface of the control nucleus first wrinkles
   into a buckyball‐like structure, followed by a transition from the
   buckyball‐like morphology to a labyrinthine pattern with deep
   invaginations along the nuclear surface (Figure [138]3A; Movie [139]S3,
   Supporting Information). As substrate stiffness decreased, the
   buckyball‐to‐labyrinth transition was delayed at characteristic times
   (t/t [0]) of 45, 90, and 105 for stiff, medium, and soft substrates,
   respectively (Figure [140]3B), as measured by the relative volume
   change (Δv/v [0], Figure [141]3C). The nuclear membrane surface tension
   (F [memb]) quantitatively assesses the substrate‐stiffness‐dependent
   differential nuclear deformation, with the occurrence of the maximum
   value delayed by reducing substrate stiffness at t/t [0] values of 30,
   75, and 100 for stiff, medium, and soft substrates, respectively
   (Figure [142]3D). These results indicate that the shrinkage of the
   nuclear shell volume and the elevation of nuclear membrane tension are
   delayed by reducing substrate stiffness, consistent with our live‐cell
   monitoring, which showed that the nucleus on stiff substrates favors
   nuclear wrinkling in response to progerin expression (Figure [143]1).
   Since we used an internal pressure to generate membrane tension (inset,
   Figure [144]3B), a lower volume modulus (K [V]), corresponding to a
   smaller ability to resist volume change, induces lower membrane
   tension. This low membrane tension, i.e., low internal pressure,
   accelerated nuclear wrinkling (Figure [145]3E; Movie [146]S3,
   Supporting Information), confirms that a greater net external pressure
   leads to faster and deeper wrinkling.
   These results demonstrate that reduced nuclear membrane tension in
   cells placed on soft substrates delays the occurrence of nuclear
   wrinkling. They further suggest that nuclear wrinkling, induced by the
   localized accumulation of permanently farnesylated prelamin A along the
   nuclear membrane, could disrupt the nuclear–cytoskeletal connection
   that is highly sensitive to changes in substrate stiffness.
2.4. Mechanosensitive Progerin Expression Induces Temporal Alterations in
Gene Profiling
   Nuclear deformation is not merely a phenotypic signature of disease but
   is strongly indicative of nuclear force‐dependent alteration of genome
   architecture and dysregulation of gene expression.^[ [147]^41 ^] Since
   progerin‐induced nuclear wrinkling is attributed to reduced NE tension
   and this biochemical response is tightly regulated by mechanosensing of
   substrate stiffness (Figures [148]1, [149]2, [150]3), we investigated
   whether progerin‐induced temporal differential NE wrinkling could also
   alter gene expression profiles in response to changes in substrate
   stiffness.
   To systematically analyze gene expression over time, we performed RNA
   sequencing for cells cultured on either a stiff glass substrate or a
   soft PAG substrate at 1.37 kPa, followed by doxycycline treatment at 6
   h intervals for up to 36 h (Figure [151]4 ). Hierarchical clustering of
   gene expression, determined by pairwise complete‐linkage clustering
   analysis, revealed that the tested samples were primarily grouped by
   doxycycline treatment time, irrespective of substrate stiffness (i.e.,
   0, 6, 12, 18, and 36 h). However, only the samples treated with
   doxycycline for 24 and 30 h formed distinct groups based on substrate
   stiffness, i.e., soft 1.37 kPa substrates at 24 and 30 h versus stiff
   glass substrates at the same time points (Figure [152]4A; Figure
   [153]S5A, Supporting Information). Further analysis of the number of
   differentially expressed genes (DEGs) in experimental conditions
   compared to the doxycycline‐untreated control group (denoted as glass 0
   h or 1.37 kPa 0 h) indicated that the largest changes in gene
   expression levels, including both upregulation and downregulation,
   occurred at 24 and 30 h after doxycycline treatment for each substrate
   stiffness (Figure [154]4B). Interestingly, these time points coincided
   with the reduction of NE tension and the formation of nuclear wrinkles,
   occurring at 24 and 25 h on stiff substrates, or 30 and 32 h on soft
   substrates, respectively (Figures [155]1 and [156]2). These results
   imply that mechanosensitive progerin expression leads to differential
   changes in gene expression as NE tension‐dependent nuclear wrinkling
   progresses.
Figure 4.
   Figure 4
   [157]Open in a new tab
   Time‐dependent alteration of gene profiling and signaling pathways in
   response to mechanosensitive progerin expression. A–D)
   Progerin‐induced, time‐dependent differential gene expression in
   response to changes in substrate stiffness (stiff glass substrates vs
   soft PAG substrates of 1.37 kPa). The similarity in gene expression
   profiles was assessed by Euclidean distance and complete linkage
   clustering, where the height of the dendrogram represents the Euclidean
   distance between clusters, indicating the similarity in expression
   profiles (A). Note that gene expression profiles remained similar
   between cells on stiff glass and soft PAG substrates at 18 h but
   clustered by substrate stiffness at 24 and 30 h (marked by red boxes),
   and reclustered at 36 h. The number of differentially regulated genes
   is displayed for each substrate stiffness by comparing with the
   doxycycline‐untreated control condition (glass 0 h on the left, 1.37
   kPa 0 h on the right), where yellow and blue bars indicate upregulated
   and downregulated genes, respectively (B). Volcano plots display
   doxycycline treatment time‐dependent evolution of log2 fold changes in
   LMNA and LINC complex‐associated genes (e.g., SUN1, SUN2, SYNE1, SYNE2,
   and SYNE3) on stiff glass substrates (top row) or soft PAG substrates
   (bottom row) (C). Note that while LMNA expression increased from 6 h
   after doxycycline treatment on both stiff glass and soft PAG
   substrates, significant increases in LMNA, SYNE2, and SYNE3 were
   observed at 12 h on stiff substrates but at 18 h on soft substrates
   (marked by red dotted boxes), indicating delayed expression of nesprin
   on soft substrates. Gene ontology (GO) analysis was performed comparing
   the control glass and soft PAG substrates at different doxycycline
   treatment times (12, 18, 24, 30, and 36 h) for cellular components,
   biological processes, and molecular functions, with term sizes between
   10 and 500 (D). In panel B, the criteria for significant changes in
   gene expression were fold change ≥ |2| and raw p‐value < 0.05. In panel
   C, yellow and red dots indicate specific gene expression levels
   corresponding to fold change ≥ |1.5|, raw p‐value < 0.5, and fold
   change ≤ |1.5| with raw p‐value < 0.5, respectively, with LINC
   complex‐associated genes colored blue. In panel D, adjusted p‐values
   reported from g:Profiler were derived using a one‐sided hypergeometric
   test and corrected by the Benjamini–Hochberg method (***: p < 0.001,
   **: p < 0.01, *: p < 0.05). E–H) Heatmap analysis of GO terms related
   to mechanosensing of substrate stiffness. Representative signaling
   pathways, including the Notch signaling pathway (GO:0007219, E), BMP
   signaling pathway (GO:0030509, F), extracellular structure organization
   (GO:0043062, G), and tissue homeostasis (GO:0001894, H), were
   visualized. Euclidean distance was used as the distance metric, and
   complete linkage was applied for hierarchical clustering in the
   analysis of each dataset. For further details, refer to the
   Experimental Section.
   Since the LINC complex bridging between the extra‐nuclear–cytoskeletal
   network and nuclear membrane‐associated proteins that physically
   interact with chromosomal architecture regulates the transmission of
   biophysical stimuli into the nucleus,^[ [158]^21 , [159]^41 ^] we
   conducted RNA sequencing to investigate whether progerin expression
   regulates the expression of LINC complex‐associated genes encoding SUN
   and nesprin proteins. Compared to the doxycycline‐untreated control
   conditions, our results showed a significant increase in the expression
   of LMNA, SYNE2, and SYNE3 genes in response to increasing doxycycline
   treatment across substrate stiffness, where we also observed
   doxycycline treatment time‐dependent differential expression of these
   genes (Figure [160]4C). For instance, LMNA gene expression was observed
   at 6 h in both stiff and soft substrates, whereas SYNE2 and SYNE3
   expression increased after 12 h of doxycycline treatment on stiff
   substrates but after 18 h on soft substrates (Figure [161]4C). These
   findings align with previous results showing that doxycycline‐induced
   progerin expression (Figure [162]1), reduction of nuclear tension
   (Figure [163]2), and nuclear wrinkling (Figures [164]1, [165]2, [166]3)
   were delayed on soft substrates compared to cells on stiff substrates.
   To further investigate whether substrate stiffness‐dependent
   differential gene expression induced by doxycycline‐inducible progerin
   expression could functionally alter mechanosensation‐mediated pathways,
   we used the gene set enrichment analysis (GSEA) database (Figure
   [167]S5B–E, Supporting Information). Specifically focusing on
   actomyosin‐related genes (GSEA C5>GO, 286 genes) and nuclear
   membrane‐related genes (GSEA C5>GO, 417 genes) obtained after 24 h of
   doxycycline treatment, 13 actomyosin‐related genes and 17 nuclear
   membrane‐related genes were identified as DEGs (Figure [168]S5B–D,
   Supporting Information). Furthermore, the changes in the expression of
   these DEGs occurred 24 h after doxycycline treatment on stiff
   substrates and 30 h after doxycycline treatment on soft substrates
   (Figure [169]S5E, Supporting Information).
   Among the top 20 GO enrichment analysis results, based on adjusted
   p‐values, our investigation of cellular components, biological
   processes, and molecular functions identified a significant delay in
   soft substrates compared to that in stiff substrates (Figure [170]4D).
   These results were compared with RNA sequencing data from the gene
   expression omnibus public datasets [171]GSE141950 and [172]GSE118633,
   which analyzed dermal fibroblasts from healthy individuals and patients
   with HGPS (Figure [173]S6, Supporting Information). [174]GSE141950
   analysis revealed distinct gene expression patterns in fibroblasts from
   patients with HGPS (Figure [175]6A,B, Supporting Information).
   Furthermore, pathway enrichment analysis of cellular components,
   biological processes, and molecular functions revealed similar changes
   in these pathways (Figure [176]4D; Figure [177]S6C, Supporting
   Information). Moreover, [178]GSE118633 analysis revealed distinct gene
   expression patterns in fibroblasts from patients with HGPS, confirming
   changes in the same pathways (Figure [179]4D; Figure [180]S6D–F,
   Supporting Information). These results not only confirm that our
   developed Tet‐On‐inducible progerin‐expressing HeLa cells effectively
   model HGPS but also demonstrate that substrate stiffness‐dependent
   differential progerin expression could modify the onset of gene
   expression that regulates multiple signaling pathways.
   Finally, we evaluated whether mechanosensitive progerin expression
   could regulate Notch signaling and bone morphogenetic protein (BMP)
   signaling. The Notch signaling pathway, highly sensitive to mechanical
   signals, regulates cell and tissue fate in most tissues,^[ [181]^42 ^]
   and the BMP signaling pathway, involving nuclear membrane proteins, is
   directly regulated by mechanical signal transduction pathways without
   autocrine ligands, occurring at the receptor, cytoplasmic, and nuclear
   levels.^[ [182]^43 ^] Heatmap analysis confirmed that Notch and BMP
   signaling were altered by doxycycline‐induced progerin expression in a
   substrate stiffness‐dependent manner (Figure [183]4E,F). Furthermore,
   we observed a distinct regulation of gene expression involved in
   extracellular structure organization in response to changes in
   substrate stiffness (Figure [184]4G), which is necessary for tissue
   homeostasis.^[ [185]^44 ^] Heatmap analysis of tissue
   homeostasis‐regulating differential gene expression further confirmed
   that mechanosensitive progerin expression could functionally regulate
   tissue homeostasis (Figure [186]4H).
   Together, these results suggest that substrate stiffness‐dependent
   differential progerin expression regulates gene expression, ultimately
   altering mechanosensitive signaling pathways.
2.5. Mechanosensitive Progerin Expression Modulates the Spatiotemporal
Reorganization of Heterochromatin
   Progerin exhibits a strong binding affinity for histone‐lysine
   N‐methyltransferase SUV39H1, preventing its proteasomal degradation and
   thereby increasing epigenetic modifications to the DNA packaging
   protein histones, e.g., H3K9me3. This, in turn, reduces DNA repair
   capacity and accelerates senescence.^[ [187]^45 ^] Meanwhile, elevated
   substrate stiffness correlates with enhanced levels of H3K9me2/3.^[
   [188]^46 ^] Thus, we investigated whether the substrate
   stiffness‐dependent mechanosensitive alteration of progerin expression
   could modulate H3K9me2/3 levels, indicative of heterochromatin
   formation and transcriptional silencing.
   To assess H3K9me2/3 protein expression levels, western blot analysis
   was performed on cells cultured on either stiff glass substrates or
   soft PAG substrates (1.37 kPa), which were treated with doxycycline at
   12 h intervals for up to 36 h (Figure [189]5A–D). Cells cultured on
   stiff glass substrates showed a significant increase in H3K9me2/3
   expression 24 h after doxycycline treatment, with levels doubling by 36
   h (Figure [190]5A,C). In contrast, cells cultured on soft PAG
   substrates showed increased H3K9me2/3 expression after 24 h of
   treatment, but the levels only slightly increased further by 36 h
   (Figure [191]5B,D). These results suggest that temporal alteration of
   histone modifications is also accompanied by the substrate
   stiffness‐dependent differential progerin expression.
Figure 5.
   Figure 5
   [192]Open in a new tab
   Time‐dependent differential epigenetic modifications in response to
   mechanosensitive progerin expression. A–D) Quantification of
   doxycycline treatment time‐dependent differential expression of
   H3K9me2/3 in response to changes in substrate stiffness. H3K9me2/3
   expression was quantified by immunoblotting against H3K9me2/3 and GAPDH
   antibodies in progerin‐expressing Tet‐On HeLa cells placed on control
   stiff substrates (glass, A) and soft PAG substrates (1.37 kPa, B) with
   doxycycline treatment every 12 h for up to 36 h. (C,D) Total protein
   expression increased with doxycycline treatment in each condition. In
   panels C and D, three independently performed experiments were averaged
   and normalized to the values in 0 h condition. Error bars indicate the
   S.E.M., and one‐way ANOVA using Tukey's test was applied for comparison
   between groups (****: p < 0.0001, ***: p < 0.005, *: p < 0.05, NS: not
   significant). E–N) Spatiotemporal alterations of heterochromatic
   histone modifications in doxycycline‐induced progerin‐expressing cells
   placed on varying substrate stiffness. Tet‐On HeLa cells expressing
   mCherry‐tagged Δ50 LMNA (red) placed on control glass (E) and PAG
   substrates of 1.37 kPa (F) were immunostained for nuclei (DAPI, blue)
   and H3K9me2/3 (green) every 12 h after doxycycline treatment.
   3D‐rendered nuclei, reconstructed from z‐stacked confocal fluorescent
   images, depict that doxycycline treatment time‐dependently increased
   progerin expression, inducing H3K9me2/3 clustering in the nuclear
   interior (E–N). H3K9me2/3 clusters were detected after 24 h on stiff
   substrates (glass, E,G–J) but appeared after 36 h on soft substrates
   (1.37 kPa, F,K–N). Yellow dotted lines indicate the nuclear boundary as
   determined by DAPI staining; white arrowheads indicate clustered
   H3K9me2/3 (E,F). Fluorescence intensity profiles monitored by line
   scanning through the maximum intensity projected nuclear images show
   that H3K9me2/3 clusters largely alternate with progerin staining. More
   intensive peaks were detected in nuclei of cells placed on stiff
   substrates compared to those on soft substrates (G–J vs K–N), where red
   and green arrowheads indicate fluorescence intensity peaks
   corresponding to progerin and H3K9me2/3 expression, respectively.
   Given that spatially resolved epigenetic modifications of histones can
   induce specific alterations in chromatin accessibility and
   transcriptional phenotypes,^[ [193]^46 ^] we also investigated the
   intranuclear distribution of H3K9me2/3 (Figure [194]5E–N).
   Cross‐sectional analysis of 3D‐reconstructed confocal images of nuclei
   costained for progerin and H3K9me2/3 revealed that increased H3K9me2/3
   levels were associated with clustering of H3K9me2/3 within the nucleus
   24 h after doxycycline treatment on stiff substrates, where H3K9me2/3
   clustering was alternatively stained for progerin expression
   (Figure [195]5E). Fluorescence intensity profiles obtained by line
   scanning through the nuclear interior confirmed that progerin
   expression and H3K9me2/3 clustering displayed distinct alternative
   staining patterns at 24 h, with fluorescence profiles intensifying by
   36 h (Figure [196]5G–J). However, on soft PAG substrates, intranuclear
   H3K9me2/3 clustering predominated after 36 h of doxycycline treatment,
   coinciding with enhanced progerin expression at the same time
   (Figure [197]5F). Line scanning through the nuclear interior revealed
   distinct alternative staining for H3K9me2/3 clusters and progerin at 36
   h (Figure [198]5K–N).
   These results indicate that substrate stiffness‐dependent differential
   onset of progerin expression leads to a time‐dependent progression of
   heterochromatin levels, accompanied by clustering of H3K9me2/3 within
   the nucleus. This strongly suggests that the delayed gene expression in
   cells placed on soft substrates, which transmit reduced forces to the
   nucleus, can be attributed to incomplete heterochromatin structure.
2.6. Substrate Stiffness‐Dependent Differential Progerin Expression Is
Mediated by Distinct Transcription Factor Binding Motifs
   Alterations in gene expression and epigenetic modifications are
   regulated by substrate stiffness‐dependent differential progerin
   expression (Figures [199]4 and [200]5). In line with previous reports
   indicating that lamin A depletion enhances chromatin mobility,^[
   [201]^47 ^] these results strongly imply that progerin‐induced
   differential gene expression patterns in response to changes in
   substrate stiffness can be attributed to variations in transcription
   factor binding motifs.
   To test this hypothesis, we first examined whether progerin expression
   alters internuclear chromatin dynamics by comparing telomere motion in
   dermal fibroblasts derived from a 3‐year‐old healthy individual and a
   same‐aged patient with HGPS, denoted as 3YR (control) and 3YR (HGPS),
   respectively (Figure [202]6A–C). Time‐lapse monitoring of GFP‐labeled
   telomeres bound by the telomeric repeat‐binding factor 2 (TRF2)
   revealed enhanced telomere movement in HGPS cells compared to control
   cells (Figure [203]6A,B). Single‐particle tracking, followed by
   calculation of the mean squared displacement (MSD) for the recorded
   trajectories, confirmed more diffusible chromatin motion in HGPS cells
   than in control cells, representing a confined motion (Figure [204]6C;
   Movie [205]S4, Supporting Information).
Figure 6.
   Figure 6
   [206]Open in a new tab
   Differential chromatin accessibility in response to mechanosensitive
   progerin expression. A–C) Differential chromatin mobility in response
   to progerin expression. Time‐lapse tracking of fluorescence‐tagged
   chromatin was performed in TRF2 (telomeric repeat‐binding factor
   2)‐transfected human dermal fibroblasts obtained from a three‐year‐old
   healthy control (denoted as 3 YR (Control), A) and an HGPS patient
   (denoted as 3 YR (HGPS), B), where nine randomly selected chromatin
   trajectories are displayed. (C) Quantitative analysis of the mean
   squared displacement (MSD) at each time lag indicates enhanced
   chromatin mobility in HGPS patients compared to the healthy control.
   D–R) ATAC sequencing‐based identification of differential key
   transcription factor (TF) binding motifs in response to substrate
   stiffness‐dependent progerin expression. Representative de novo TF
   binding motifs in Tet‐On HeLa cells expressing progerin were identified
   between cells on control stiff substrates (denoted as glass, D) and
   cells on soft PAG substrates (denoted as 1.37 kPa, E) at 12 h intervals
   after doxycycline treatment for 36 h using Homer software. The bar
   graph indicates the percentage of target binding motifs for CTCF,
   FOS::JUNB, TEAD family, KLF1, and ZNF331 (F), where each bar represents
   fold enrichment, defined as the percentage of target sequences with the
   motif divided by the percentage of background sequences with the motif.
   Data are normalized to doxycycline‐untreated control groups (denoted as
   glass 0 h, 1.37 kPa 0 h). Doxycycline treatment increased fold
   enrichment for TEAD (G,H), CTCF (K,L), and NFkB‐p65‐Rel (O,P), but
   decreased fold enrichment for JunB (I,J) and KLF1 (M,N) in control
   stiff substrates, with these changes diminished in soft PAG substrates.
   ATAC‐seq tracking of ZNF331 (Q), a known transcriptional repressor, and
   BMP2 (R), a component of mechanosensory pathways, was visualized using
   Integrative Genomics Viewer (IGV), where blue and red peaks indicate
   stiff glass and soft PAG substrates, respectively.
   Because these results strongly imply that nuclear wrinkling, typically
   observed in patients with HGPS, is associated with enhanced chromatin
   dynamics due to progerin‐induced loss of mechanical integrity along the
   nuclear lamina,^[ [207]^47 , [208]^48 ^] we further investigated
   whether progerin‐induced differential chromatin dynamics regulate
   genome‐wide chromatin accessibility. Specifically, we performed a
   high‐throughput assay for transposase‐accessible chromatin using
   sequencing (ATAC‐seq) to assess whether progerin‐induced distinct,
   time‐dependent gene expression patterns in response to changes in
   substrate stiffness are attributable to variations in transcription
   factor binding motifs. The calculation of transcription start site
   (TSS) enrichment scores revealed that chromatin‐accessible regions are
   enriched at TSSs, and the distribution of aligned fragment lengths
   obtained from all tested samples confirmed the high quality of ATAC‐seq
   data (Figure [209]S7A,B, Supporting Information). We observed that
   chromatin‐accessible regions were similarly distributed across the
   genome under different experimental conditions (Figure [210]S7C,D,
   Supporting Information). Analysis of transcription factor (TF) binding
   motifs within chromatin‐accessible regions revealed specific changes in
   TF binding motifs in response to doxycycline‐induced progerin
   expression on stiff glass substrates (denoted as glass, Figure [211]6D)
   and soft PAG substrates (denoted as 1.37 kPa, Figure [212]6E), selected
   from the top ten TFs predicted to have binding motifs in
   chromatin‐accessible regions for each condition (Figure [213]S7E,F,
   Supporting Information).
   In particular, as represented by the percentage of targets, we noted
   that the expression of differential TF‐binding motifs for CCCTC‐binding
   factor (CTCF), FOS::JUNB, TEA domain (TEAD) family, Kruppel‐like factor
   1 (KLF1), and zinc finger protein 331 (ZNF331) was delayed on soft
   substrates compared to stiff substrates (Figure [214]6F). These results
   suggest that progerin‐induced differential gene expression
   (Figure [215]4), following enhanced H3K9me2/3 clustering
   (Figure [216]5), could be attributed to differences in TF‐binding
   motifs. Progerin‐induced alterations in heterochromatin structure could
   modulate the configuration of TF binding, ultimately regulating
   substrate stiffness‐dependent gene expression.
   A systematic comparison of fold enrichment values normalized to the
   untreated doxycycline group (0 h) identified differentially regulated
   TFs that promote multiple aging‐associated cellular mechanisms, while
   all the detected TFs were expressed at lower levels on soft substrates
   than on stiff substrates (glass vs 1.37 kPa, Figure [217]6G–P). For
   instance, the expression of TEAD (Figure [218]6G,H), CTCF
   (Figure [219]6K,L), and NFkB‐p65‐Rel (Figure [220]6O,P), regulating
   pathological processes in HGPS, including cell proliferation, chromatin
   structure, and inflammatory responses,^[ [221]^49 ^] progressively
   increased with doxycycline treatment. In contrast, the expression of
   JunB (Figure [222]6I,J) and KLF1 (Figure [223]6M,N), regulating
   cellular stress responses (e.g., oxidative stress and inflammation)^[
   [224]^50 ^] and gene transcription in erythroid differentiation,^[
   [225]^51 ^] respectively, gradually decreased with doxycycline
   treatment. Meanwhile, mechanosensitive signaling‐mediated ZNF331, a
   transcriptional repressor containing the Kruppel‐associated box (KRAB)
   domain,^[ [226]^52 ^] exhibited an increase in ATAC‐seq peak height
   following doxycycline treatment, indicating enhanced accessibility of
   the ZNF331 gene (Figure [227]6Q). In contrast, the gene accessibility
   of BMP2, involved in the BMP signaling pathway, one of the key pathways
   identified in mechanosensation (Figure [228]4F), increased
   progressively by doxycycline‐induced progerin expression
   (Figure [229]6R).
   These findings suggest that nuclear wrinkling induced by progerin
   expression elevates chromatin dynamics, enhances heterochromatin
   clustering, and ultimately regulates chromatin accessibility.
2.7. Progerin‐Induced Nuclear Deformation Is Mediated by the Remodeling of
LINC Complex‐Dependent Molecular Connections with LMNA
   Lamin A interacts with the cytoskeleton via the LINC complex, which is
   composed of SUN proteins in the inner nuclear membrane and nesprin
   isoforms, the cytoplasmic domains of KASH proteins in the outer nuclear
   membrane, enabling the transmission of cytoskeletal forces to the
   nuclear membrane.^[ [230]^53 ^] We previously demonstrated that
   progerin expression reduced nuclear tension by incorporating the
   nesprin tension sensor, mimicking nesprin 2 bridging between the actin
   cytoskeleton and SUN proteins (Figure [231]2). Based on this, we
   hypothesized that progerin‐induced nuclear deformation could be
   mediated by the differential interactions between the LINC complex and
   LMNA.
   To determine whether the evolution of NE wrinkling in response to
   progerin expression was associated with chromosomal interactions
   between DNA and LINC proteins, we first analyzed chromatin
   accessibility of LMNA, SUN1, and SYNE2 using ATAC‐seq (Figure
   [232]S8A–F, Supporting Information). After doxycycline treatment, these
   genes progressively enhanced chromatin accessibility on both stiff and
   soft substrates (Figure [233]7A–F). While LMNA gene accessibility
   increased on both substrate types 12 h post‐doxycycline treatment,
   however, the accessibility of SUN1 and SYNE2 genes was delayed on soft
   substrates, with a similar increase observed on both stiff and soft
   substrates at 36 h (Figure [234]7A–F; Figure [235]S8A–F, Supporting
   Information). Consistent with hierarchical clustering analysis of
   RNA‐seq indicating that samples were clustered by substrate stiffness
   at 24 and 30 h, but clustered by doxycycline treatment time at 36 h
   (Figure [236]4A), these results suggest that substrate
   stiffness‐dependent differential chromatin accessibility was dominant
   at 24 h of progerin expression, but diminished at 36 h of doxycycline
   exposure. Together with the previous result exhibiting a delayed
   reduction of NE tension on soft substrates (Figure [237]2), these
   findings suggest that Tet‐On‐inducible progerin expression modulates
   LMNA gene expression, simultaneously upregulating SUN1 and SYNE2, while
   reduced NE tension induced by progerin expression facilitates LINC
   complex‐mediated molecular binding.
Figure 7.
   Figure 7
   [238]Open in a new tab
   LINC complex‐mediated remodeling of LMNA‐associated nuclear tethering
   in progerin‐expressing cells. A–F) ATAC‐seq‐based identification of
   chromatin accessibility for LMNA‐associated LINC complex components in
   progerin‐expressing cells. ATAC‐seq peaks for LMNA (A,B), SUN1 (C,D),
   and SYNE2 (E,F) in progerin expression‐induced cells treated with
   doxycycline at 12 h intervals for up to 36 h on control glass (A,C,E)
   and soft PAG substrates (B,D,F) were visualized by Integrative Genomics
   Viewer (IGV). Note that as doxycycline treatment progresses, peaks for
   LMNA, SUN1, and SYNE2 in both control glass and soft PAG increase,
   indicating enhanced chromatin accessibility, but the peak height in
   soft substrates remains lower than that in control glass, approaching a
   similar level at 36 h. G–N) Differential expression of NE‐associated
   proteins and nuclear wrinkling in response to progerin expression.
   Tet‐On HeLa cells expressing mCherry‐tagged LMNA or Δ50 LMNA (progerin)
   were immunostained for SUN1 (green) and nucleus (DAPI, blue) (G,H) or
   nesprin 2 (green) and nucleus (DAPI, blue) (I,J) before (−Dox) and
   after (+Dox) doxycycline treatment. While doxycycline‐induced
   expression of mCherry‐tagged LMNA did not alter the SUN1 (G,K) and
   nesprin 2 contents (I,L), doxycycline‐induced expression of
   mCherry‐tagged Δ50 LMNA significantly increased SUN1 (H,K) and nesprin
   2 (J,L). Compared to doxycycline‐induced LMNA expression, which did not
   induce changes in nuclear shape (G,I,M), Δ50 LMNA expression
   significantly increased NE wrinkling (H,J,M). SUN1 and nesprin 2
   expression was more sensitive to Δ50 LMNA expression than to LMNA
   expression (N). In panels K–N, >50 nuclei were analyzed for each
   condition; error bars indicate the standard error of the mean (S.E.M.);
   and Student's t‐test was applied for comparison between two groups
   (****: p < 0.0001, ***: p < 0.001, NS: not significant). O–T)
   Progerin‐induced differential interaction in LMNA‐associated LINC
   proteins. The strength of molecular interaction between SUN1 and
   nesprin 2 (O,Q,R) or between SUN1 and LMNA (P,S,T) was estimated by
   quantifying the number and total intensity of proximity ligation assay
   (PLA) signals (red dots) in DAPI‐stained nuclei (blue). Hemispherical
   and cross‐sectional views of 3D‐rendered nuclei showed that punctate
   PLA signals were preferentially localized along the nuclear periphery.
   Note that doxycycline‐induced progerin expression significantly
   increased the PLA signals of SUN1 associated with LMNA (S,T), while PLA
   signals of SUN1 associated with nesprin 2 remained unchanged (Q,R). In
   panels Q, R, S, and T, >50 nuclei were analyzed for each condition;
   error bars indicate the S.E.M.; and an unpaired t‐test was applied to
   assess statistical significance (***: p < 0.001, NS: not significant).
   To test this notion, we examined the expression levels of SUN1 and its
   binding partner, nesprin 2, in response to progerin expression by
   comparing mCherry‐tagged LMNA and progerin (denoted as Δ50
   LMNA)‐expressing Tet‐On HeLa cells (Figure [239]7G–J). Quantitative
   immunofluorescence microscopy revealed that both SUN1 and nesprin 2
   were upregulated in progerin‐expressing cells but not in
   LMNA‐overexpressing cells (Figure [240]7K,L), which was confirmed by
   immunoblotting analysis of total protein levels (Figure [241]S8G–J,
   Supporting Information). These results demonstrate that the
   upregulation of LINC proteins is specifically induced by progerin
   expression and not by the accumulation of intact LMNA. To assess
   whether LINC complex proteins regulate progerin‐induced nuclear
   deformation, we compared changes in NE wrinkling before and after
   doxycycline‐induced expression of LMNA or progerin. As expected, NE
   wrinkling remained unchanged before doxycycline treatment and was not
   altered by LMNA expression (left two bars, Figure [242]7M). However,
   progerin expression resulted in a sixfold increase in NE wrinkling
   (right two bars, Figure [243]7M). Moreover, we confirmed that LINC
   complex proteins were more tightly regulated by progerin expression
   than by intact LMNA expression, i.e., the expression levels of SUN1 and
   nesprin 2 increased proportionally with progerin expression but not
   with LMNA expression (Figure [244]7N). These results support the
   association between NE wrinkling and the upregulation of LINC complex
   proteins in response to progerin expression.
   Building on previous reports that SUN1 and SUN2 are not functionally
   equivalent for nuclear connection to the actin cytoskeleton,^[ [245]^54
   ^] and that SUN1, but not SUN2, is upregulated in HGPS cells,^[
   [246]^28 ^] we investigated whether doxycycline‐activated progerin
   expression could remodel the molecular connectivity of LINC‐associated
   proteins to the nucleus via upregulation of SUN1. We assessed whether
   SUN1 overexpression altered LMNA‐associated molecular interactions in
   the nuclear lamina, leading to progerin‐mediated nuclear deformation.
   To quantitatively analyze the interaction between SUN1 and LMNA or
   nesprin 2 in response to progerin expression, a proximity ligation
   assay (PLA) was performed (Figure [247]7O–T). As expected, a majority
   of punctate PLA signals were preferentially localized along the nuclear
   periphery, where LINC components form a molecular assembly with lamin
   proteins (Figure [248]7O,P). By counting the number and total intensity
   of fluorescent dots representing molecular interactions, we observed
   that doxycycline‐induced progerin expression significantly increased
   the association between SUN1 and LMNA (Figure [249]7P,S,T), whereas the
   interaction between SUN1 and nesprin 2 remained unchanged
   (Figure [250]7O,Q,R).
   Together with the upregulation of SUN1 in progerin‐expressing cells,
   the preservation of identical molecular interactions between SUN1 and
   nesprin 2 further implies that progerin accumulation in intact LMNA
   enhances the spatial proximity of SUN1 to LMNA, strongly suggesting
   that progerin expression induces nuclear deformation through the
   remodeling of SUN1‐dependent molecular connections with LMNA.
2.8. Actomyosin Contractility Regulates Nuclear Deformation by Altering
Nuclear Tension in Progerin‐Expressing Cells
   We previously demonstrated that progerin‐induced NE wrinkling was
   highly mechanosensitive (Figure [251]1) and that substrate
   stiffness‐dependent cytoskeletal tension mediated nuclear force at the
   nucleus–cytoskeletal interface (Figure [252]2). In conjunction with
   recent studies showing increased F‐actin polymerization in
   LMNA‐depleted human retinal pigment epithelial cells^[ [253]^29 ^] and
   elevated RhoA activation in Z24^−/− MSCs,^[ [254]^30 ^] we hypothesized
   that progerin‐induced nuclear deformation could be regulated by
   alterations in nuclear tension driven by actomyosin contractility.
   To investigate whether doxycycline‐controlled progerin expression could
   alter cytoskeletal tension, we examined F‐actin content and myosin
   activity using quantitative immunofluorescence microscopy (Figure
   [255]8A–C). As predicted, progerin‐expressing cells exhibited increased
   F‐actin organization (Figure [256]8A,B). Furthermore, we observed a
   significant increase in myosin II content relative to F‐actin, assessed
   by measuring the intensity of phospho‐myosin light chain 2 (pMLC2)
   normalized to individual F‐actin fibers (Figure [257]8C). These
   findings confirm that myosin‐dependent cytoskeletal tension is elevated
   in progerin‐expressing cells.
Figure 8.
   Figure 8
   [258]Open in a new tab
   Actomyosin contractility‐dependent nuclear deformation in
   progerin‐expressing cells. A–C) Differential actomyosin contractility
   in response to doxycycline‐induced progerin expression.
   Doxycycline‐inducible progerin‐expressing HeLa cells were immunostained
   for F‐actin (green), phospho‐myosin light chain 2 (pMLC2, red), and
   nucleus (DAPI, blue) before (−Dox) and after (+Dox) doxycycline
   treatment (A). (Insets) The details of pMLC2 staining along the actin
   stress fibers. Doxycycline‐induced progerin expression significantly
   increased the F‐actin (B) and pMLC2 (C) contents, which were normalized
   to cell area and actin stress fibers, respectively. In panels B and C,
   >60 cells were analyzed for each condition; error bars indicate the
   S.E.M.; unpaired t‐test was applied (***: p < 0.001). D–F) Differential
   formation of F‐actin and pMLC2 in doxycycline‐controlled
   progerin‐expressing Tet‐On HeLa cells in response to pharmaceutical
   inhibition of myosin‐dependent cytoskeletal tension. Cells were
   immunostained for nuclei (DAPI, blue), F‐actin (green), and pMLC2 (red)
   before (−Dox) and after (+Dox) doxycycline treatment, where
   differential concentrations of myosin‐II inhibiting blebbistatin were
   added (D). Compared to −Dox control cells, doxycycline‐induced
   progerin‐expressing cells showed significantly enhanced F‐actin, which
   remained unchanged in response to specific disruption of myosin
   activity (E). Significantly increased pMLC2 content due to
   doxycycline‐induced progerin expression was reversed by increasing the
   concentration of blebbistatin (F). Treating doxycycline‐induced
   progerin‐expressing cells with 15 µm blebbistatin fully restored their
   pMLC2 content to the level of doxycycline‐untreated progerin
   nonexpressing cells (F). In panels E and F, >150 cells were analyzed
   per condition; error bars indicate the S.E.M.; one‐way ANOVA using
   Tukey's test was applied (***: p < 0.001, **: p < 0.05, NS: not
   significant). G,H) Actomyosin contractility‐dependent differential
   changes of nuclear tension. Representative nesprin tension sensor‐based
   FRET signals along the nuclear membrane of doxycycline‐inducible
   progerin‐expressing cells were captured before (−Dox) and after (+Dox)
   doxycycline treatment in the presence of DMSO and 10 or 15 µm
   blebbistatin (G). Compared to doxycycline‐untreated control,
   doxycycline‐induced progerin expression significantly enhanced the FRET
   ratio, which was gradually diminished by increasing the concentration
   of blebbistatin and fully restored to the level of
   doxycycline‐untreated control condition by 15 µm blebbistatin treatment
   (H). I–K) Tight regulation of SUN1 expression and NE wrinkling in
   response to changes in actomyosin contractility. Tet‐On HeLa cells
   expressing mCherry‐tagged Δ50 LMNA (progerin) were treated with DMSO
   and 10 or 15 µm blebbistatin in the absence (−Dox) and presence (+Dox)
   of doxycycline before immunostaining for nucleus (DAPI, blue) and SUN1
   (green) (I). Compared to the doxycycline‐untreated control, SUN1
   expression and NE wrinkling were significantly increased in
   doxycycline‐treated progerin‐expressing cells, which was gradually
   diminished by increasing the concentration of blebbistatin and fully
   restored to the level of the control condition by 15 µm blebbistatin
   treatment (J,K). In panels H, J, and K, >20 cells were analyzed per
   condition; error bars indicate the S.E.M.; one‐way ANOVA using Tukey's
   test was applied for comparison between groups (***: p < 0.001, **: p <
   0.05, NS: not significant).
   Next, to determine whether actomyosin contractility regulates nuclear
   tension‐mediated NE wrinkling, we treated progerin‐expressing cells
   with the myosin II inhibitor blebbistatin at varying concentrations
   (Figure [259]8D). Consistent with previous results (Figure [260]8A–C),
   doxycycline‐induced progerin expression significantly increased F‐actin
   and pMLC2 intensities in dimethyl sulfoxide (DMSO)‐treated control
   cells (Figure [261]8E,F). However, pMLC2 content gradually decreased
   with increasing drug concentration, and 15 µm blebbistatin treatment
   fully restored pMLC2 content to the level observed in
   doxycycline‐untreated progerin nonexpressing cells (Figure [262]8F),
   while F‐actin content remained unchanged (Figure [263]8E).
   We assessed actomyosin contractility‐dependent nuclear tension using
   the nesprin tension sensor in doxycycline‐induced progerin‐expressing
   cells treated with varying concentrations of blebbistatin
   (Figure [264]8G). Surprisingly, we found that a gradual reduction in
   myosin activity enhanced nuclear tension, as indicated by the FRET
   ratio (Figure [265]8H). Treatment of doxycycline‐induced
   progerin‐expressing cells with 15 µm blebbistatin reduced pMLC2
   expression to levels observed in doxycycline‐untreated progerin
   nonexpressing cells (Figure [266]8F) without disrupting the F‐actin
   content (Figure [267]8E). Under these conditions, the FRET ratio was
   fully restored (Figure [268]8H).
   We further confirmed that enhanced nuclear tension (i.e., decreased
   FRET ratio) in response to pharmaceutical inhibition of pMLC2
   expression suppressed SUN1 expression in progerin‐expressing cells
   (Figure [269]8I,J). Consequently, the level of NE wrinkling in 15 µm
   blebbistatin‐treated progerin‐expressing cells was fully restored to
   that observed in doxycycline‐untreated cells (Figure [270]8K). These
   data suggest that inhibition of myosin‐dependent cytoskeletal tension
   reverses nuclear deformation by downregulating SUN1 expression and
   restoring NE tension.
   Taken together, our results demonstrate that actomyosin contractility
   regulates nuclear deformation through alterations in nuclear tension in
   progerin‐expressing cells.
2.9. Inhibition of SUN1 Recovers Progerin‐Induced Nuclear Wrinkling in HGPS
Cells
   Upregulated SUN1 expression in progerin‐expressing cells resulted in
   the remodeling of molecular connections with LMNA (Figure [271]7), and
   progerin‐induced nuclear deformation was regulated by pMLC2‐dependent
   nuclear tension (Figure [272]8). Together with a previous study showing
   that SUN1 depletion reduces actomyosin activity without disrupting the
   expression of nesprin 2 in vascular smooth muscle cells,^[ [273]^17a ^]
   we hypothesized that the level of SUN1 expression in
   progerin‐expressing cells could determine myosin‐dependent cytoskeletal
   tension, ultimately regulating nuclear tension‐dependent nuclear
   deformation.
   To directly assess the causality between SUN1 expression and myosin
   activity, we inhibited SUN1 expression by transfecting
   progerin‐expressing Tet‐On HeLa cells with small interfering RNA
   (siRNA) and compared them to control siRNA‐transfected cells (Figure
   [274]9A). Consistent with previous data (Figures [275]7K and [276]8C),
   doxycycline‐induced progerin‐expressing cells transfected with control
   siRNA showed significantly increased levels of SUN1 and pMLC2, whereas
   siRNA‐mediated SUN1 depletion reduced these levels comparable to those
   observed in doxycycline‐untreated, progerin‐negative cells
   (Figure [277]9B,C). Confirming that the changes in pMLC2 expression in
   Tet‐On HeLa cells in response to doxycycline‐induced progerin
   expression and/or siRNA transfection were largely proportional to
   changes in SUN1 expression (Figure [278]9A–C), a strong correlation
   between SUN1 and pMLC2 expression was detected, regardless of specific
   conditions (Figure [279]S9, Supporting Information). These results
   indicate that increased pMLC2 expression in progerin‐expressing cells
   is induced by SUN1 upregulation.
Figure 9.
   Figure 9
   [280]Open in a new tab
   SUN1‐mediated nuclear tension regulates progerin‐induced nuclear
   deformation. A–C) SUN1‐mediated modulation of actomyosin activity in
   progerin‐expressing cells. Tet‐On HeLa cells expressing progerin were
   immunostained for F‐actin (green), pMLC2 (red), SUN1 (orange), and
   nucleus (DAPI, blue) in the doxycycline‐untreated control condition
   (−Dox) and doxycycline‐treated conditions (+Dox) with siControl
   (+Dox/+siCon) or siSUN1 (+Dox/+siSUN1)‐mediated knockdown (A).
   Doxycycline‐induced progerin expression significantly increased the
   expression levels of SUN1 and pMLC2, which were maintained in
   siControl‐transfected cells but restored to levels similar to those
   observed in doxycycline‐untreated progerin nonexpressing cells after
   transfection with siSUN1 (B,C). In panels B and C, >150 cells were
   analyzed per condition; error bars indicate the S.E.M.; one‐way ANOVA
   using Tukey's test was applied (***: p < 0.001, NS: not significant).
   D–E) SUN1 expression‐dependent NE tension. Nesprin tension sensor‐based
   FRET signals along the nuclear membrane of doxycycline‐inducible
   progerin‐expressing cells transfected with siControl (+siCon) or siRNA
   targeting SUN1 (+siSUN1) were captured before (−Dox) and after (+Dox)
   doxycycline treatment (D). Doxycycline‐induced enhanced NE tension was
   maintained in siControl‐transfected cells but restored to the level of
   the doxycycline‐untreated control condition in siSUN1‐transfected cells
   (E). In panel E, >20 cells were analyzed per condition; error bars
   indicate the S.E.M.; one‐way ANOVA using Tukey's test was applied (***:
   p < 0.001, NS: not significant). F–H) Quantification of SUN1‐mediated
   NE wrinkling. mCherry‐tagged progerin‐expressing Tet‐On HeLa cells
   transfected with siControl (+siCon) or siRNA targeting SUN1 (+siSUN1)
   were immunostained for lamin B1 (green), SUN1 (yellow), and nuclear DNA
   (DAPI, blue) in the absence (−Dox) or presence (+Dox) of doxycycline
   (F). Doxycycline‐induced progerin expression significantly increased
   the NE wrinkling, which was maintained in siControl‐transfected cells
   but reduced to the level of the doxycycline‐untreated control condition
   in siSUN1‐transfected cells (G). The Pearson product‐moment correlation
   assessment applied to the merged dataset, including all conditions,
   showed a highly correlative relationship between SUN1 expression and NE
   wrinkling (r = 0.83) (H). In panels G and H, >50 cells were analyzed
   per condition; error bars indicate the S.E.M.; and one‐way ANOVA using
   Tukey's test was applied for comparison between groups (***: p < 0.001,
   NS: not significant). I–M) SUN1‐mediated restoration of the nuclear
   morphology of HGPS fibroblasts. Human dermal fibroblasts obtained from
   a three‐year‐old healthy control (denoted by 3 YR) and an HGPS patient
   were immunostained for F‐actin (green), pMLC2 (red), SUN1 (orange), and
   nuclei (DAPI, blue), where HGPS fibroblasts were transfected with
   siControl (HGPS/+siCon) or siSUN1 (HGPS/+siSUN1) (I). Full and empty
   arrowheads indicate the smooth and wrinkled nuclear surface,
   respectively. Compared to control fibroblasts, HGPS cells displayed a
   significantly enhanced expression of SUN1 and pMLC2, which was
   maintained in siControl‐transfected cells, but transfection with siSUN1
   restored SUN1 and pMLC2 expression to levels similar to those observed
   in control cells (J,K). Nuclear wrinkles specifically featured in HGPS
   fibroblasts and siControl‐transfected HGPS fibroblasts were recovered
   in siSUN1‐transfected cells to levels comparable to those in healthy
   controls (L). Pearson correlation analysis applied to the merged
   dataset incorporating all experimental conditions showed a strong
   correlation between SUN1 expression and pMLC2 expression (red, r =
   0.98), SUN1 expression, and NE wrinkling (blue, r = 0.99) (M). In
   panels J and K, >50 cells were analyzed per condition; in panel L, >20
   cells were analyzed per condition; error bars indicate the S.E.M.;
   one‐way ANOVA using Tukey's test was applied for comparison between
   groups (***: p < 0.001, **: p < 0.05, NS: not significant).
   As reduced nuclear tension (i.e., increased FRET ratio) was restored by
   inhibiting pMLC2 expression (Figure [281]8H), we assessed whether
   transfection with SUN1‐siRNA could also restore this reduced nuclear
   tension (Figure [282]9D). In progerin‐expressing cells transfected with
   SUN1‐siRNA (+Dox/+siSUN1), the increased FRET ratio induced by
   doxycycline‐mediated progerin expression (+Dox or +Dox/+siCon) was
   reduced to the level observed in doxycycline‐untreated cells (−Dox)
   (Figure [283]9E). This confirmed that the reduced nuclear tension in
   progerin‐expressing cells was mediated by SUN1 upregulation.
   Measurement of NE wrinkling indicated that nuclear deformation caused
   by progerin‐induced reduction in nuclear tension was reversed by
   transfection with SUN1‐siRNA (Figure [284]9F,G), indicating a strong
   correlation between SUN1 content and the level of NE wrinkling under
   all conditions (Figure [285]9H). Combined with previous data showing
   that progerin‐induced NE wrinkling is mediated by a SUN1‐dependent
   reduction in nuclear tension, accompanied by upregulated
   myosin‐associated cytoskeletal tension (Figures [286]7 and [287]8), and
   that gradual inhibition of pMLC2 expression reverses the upregulation
   of SUN1 and NE wrinkling (Figure [288]8J,K), these findings suggest
   that SUN1 upregulation is responsible for nuclear deformation in
   progerin‐expressing cells.
   To extend our findings to human patients suffering from an
   accelerated/premature aging disorder, we utilized dermal fibroblasts
   derived from a 3‐year‐old patient with HGPS and compared them to
   control cells from a healthy individual of the same age. Transfection
   of HGPS fibroblasts with SUN1–siRNA significantly reduced the
   expression of both SUN1 and pMLC2, without disrupting F‐actin
   organization, which remained unchanged over 5 days (Figure [289]S10,
   Supporting Information). By systematically comparing fibroblasts from
   healthy individuals (denoted as 3 YR), we found that both HGPS
   fibroblasts and siControl‐transfected HGPS fibroblasts exhibited
   significantly increased levels of SUN1 and pMLC2, which was reversed by
   transfection with siSUN1 (Figure [290]9I–K). Accordingly, the level of
   NE wrinkling, which is specifically featured in HGPS fibroblasts and
   siControl‐transfected HGPS fibroblasts, was restored to healthy control
   levels by siRNA‐induced depletion of SUN1 (Figure [291]9I,L). In
   patients with HGPS, a stronger correlation was observed between SUN1
   and pMLC2 intensities, as well as nuclear envelope wrinkling. Notably,
   SUN1 knockdown using siRNA reduced these levels to those observed in
   healthy controls (Figure [292]9M).
   Together with previous results showing that doxycycline‐induced
   progerin expression results in SUN1 accumulation in the nuclear lamina,
   where nuclear tension along the SUN1–nesprin 2–F‐actin connections is
   diminished by increased pMLC2 in response to progerin expression, these
   results reconfirm that defective nuclear morphology in HGPS is induced
   by reduced nuclear tension, accompanied by SUN1 upregulation‐mediated
   pMLC2 expression rather than altered F‐actin connectivity, which
   coincides with chromosomal remodeling via modification of
   heterochromatin accessibility (Figure [293]10 ).
Figure 10.
   Figure 10
   [294]Open in a new tab
   Schematic summary depicting the functional relationship between
   SUN1‐mediated nuclear tension and NE wrinkling in response to progerin
   expression. Progerin expression accumulates LINC complex proteins SUN1
   and Nesprin 2, reorganizing the actin‐binding Nesprin‐associated LINC
   complex at the nuclear envelope, and determining the biophysical
   interactions of the nuclear–cytoskeletal connection. Although the
   molecular linkages connecting SUN1, Nesprin 2, and F‐actin remain
   unchanged in response to progerin expression, nuclear tension along the
   SUN1–Nesprin 2–F‐actin connection is reduced by increased pMLC2. In
   summary, progerin‐induced morphological defects forming the surface
   wrinkling along the nuclear lamina are determined by the accumulation
   of LINC complexes proteins at the nuclear envelope and reduced nuclear
   tension accompanied by pMLC2 via the SUN1–Nesprin 2 bridge, regulating
   the expression of various genes within the nucleus. Ultimately,
   progerin‐induced nuclear wrinkling features increased chromatin
   dynamics in the heterochromatin‐rich nuclear periphery, resulting in
   the misregulation of mechanotransduction signal pathways in the HGPS
   model. Doxycycline‐induced progerin expression exhibits mechanical
   sensitivity to variations in substrate stiffness. Approximately 10%,
   25%, and 28% of delays in onsets of progerin expression, reduction of
   nuclear tension, and nuclear wrinkling, respectively, on the soft
   substrate identifies the intracellular cytoskeletal force exerted on
   the nucleus as the origin of progerin‐induced nuclear wrinkling.
3. Discussion
   Increased cytoskeletal tension, determined by its connection to the
   nuclear envelope, is crucial in regulating the mechanical forces
   transmitted to the nucleus. However, the causal relationship between
   progerin‐induced changes in cytoskeletal dynamics and nuclear
   deformation in progerin‐expressing cells remains unclear. In this
   study, we aimed to address this gap by investigating the molecular
   mechanisms underlying nuclear morphological changes in response to
   progerin expression and their contribution to chromatin reorganization
   and aberrant gene expression patterns observed in HGPS. In this study,
   we employed a doxycycline‐inducible Tet‐On system to precisely control
   progerin expression in a human HGPS cell model, enabling real‐time
   monitoring through mCherry‐tagged proteins. This system facilitates the
   spatiotemporal monitoring of nuclear deformation in response to
   mechanosensitive progerin expression while overcoming the limitation of
   primary cell availability. This approach enabled us to investigate the
   molecular mechanisms underlying progerin‐induced nuclear deformation
   and its impact on temporal chromatin remodeling and gene expression
   profiles. Moreover, this system served as an experimental platform to
   explore the relationship between progerin‐induced nuclear deformation
   and mechanosensing of substrate stiffness.
   In combination with previous reports showing that adhesion‐dependent
   cells typically display enhanced cytoskeletal tension on a rigid matrix
   compared to that on a compliant matrix to maintain mechanical
   integrity,^[ [295]^55 ^] our findings of enhanced myosin activity in
   progerin‐expressing cells indicate that the signaling pathways
   determining the progerin‐induced nuclear deformation are highly
   mechanosensitive (Figure [296]1). Previous results showed that dermal
   fibroblasts from a patient with HGPS exhibited abnormal nuclear
   morphology and even formed nuclear rupture on stiff substrates ranging
   from 10 to 20 kPa and 80 kPa, whereas such abnormalities were less
   frequently observed in soft substrates ≈3 kPa.^[ [297]^56 ^] Because
   mechanosensitive nuclear deformation is also time‐dependent
   (Figure [298]3), These results suggests that the differential onset of
   nuclear abnormalities correlates with an increase in cytoskeletal
   tension, featuring enhanced cell spreading and elongation that
   typically amplify cytoskeletal forces transmitted to the nucleus. Since
   progerin expression is regulated by cytoskeletal forces, nuclear
   responses to differential expression of progerin vary depending on the
   mechanical properties of the extracellular environment. Thus, exposure
   to substrates of varying substrate stiffness could modulate
   cytoskeletal tension and mechanotransduction, ultimately shaping the
   cellular adaptations associated with progerin‐induced phenotypes.^[
   [299]^57 ^] As varying matrix rigidity induces distinct differentiation
   of MSCs, e.g., neurogenesis on soft substrates ranging from 0.1 to 1
   kPa, and osteogenesis on stiff substrates ranging from 25 to 40 kPa,
   respectively,^[ [300]^58 ^] 1.37 kPa condition applied in our study
   closely mimics physical settings of soft tissues, whereas 34 kPa
   condition represents a microenvironment of stiffer tissues
   (Figure [301]1). This enables our system to reliably predict the
   phenotypic development associated with nuclear deformation observed in
   HGPS patients.
   We observed that progerin expression induced a critical reduction in
   nuclear tension after 24 h, followed by NE wrinkling within 1 h on
   stiff substrates, whereas these time intervals doubled on soft
   substrates (Figures [302]1 and [303]2). These results indicate that the
   synergistic increase in cytoskeletal tension can be attributed to the
   combination of progerin expression and plating of cells on the rigid
   matrix, which accelerates the individual steps of progerin‐induced
   reduction in nuclear tension and nuclear tension‐dependent NE wrinkling
   (Figures [304]1 and [305]2). The computational model system further
   demonstrated that as the mechanical resistance of the nucleus to
   external pressure increased (i.e., on a stiff substrate), the nuclear
   volume and NE tension decreased, leading to accelerated nuclear
   wrinkling compared to the nucleus on the soft substrate
   (Figure [306]3).
   Meanwhile, RNA sequencing comparing the number of DEGs following
   progerin expression on substrates of varying elastic moduli showed that
   the largest shift in gene expression levels (both upregulation and
   downregulation) occurred at 24 h after doxycycline treatment on stiff
   substrates and at 30 h on soft substrates (Figure [307]4A,B), following
   a significant reduction in NE tension (Figures [308]1 and [309]2).
   These results indicate that progerin‐induced nuclear wrinkling is not
   only mechanosensitive to changes in extracellular mechanical settings,
   mainly mediated by actomyosin contractility and nucleus–cytoskeletal
   connections,^[ [310]^16 ^] but also induces distinct pathogenetic gene
   regulation. The upregulation of LMNA, SYNE2, and SYNE3 supports this
   notion that mechanosensitive progerin expression mediates essential
   signaling pathways and functional gene regulation (Figure [311]4).
   Moreover, consistent with enhanced expression of H3K9me2/3 on rigid
   substrates,^[ [312]^46 ^] progerin‐induced epigenetic modification is
   highly mechanosensitive (Figure [313]5) and leads to enhanced chromatin
   dynamics in fibroblasts from patients with HGPS (Figure [314]6A–C).
   A comparative analysis with publicly available RNA sequencing datasets
   from primary fibroblasts of healthy individuals and patients with HGPS
   ([315]GSE141950 and [316]GSE118633) revealed distinct gene expression
   profiles in HGPS fibroblasts (Figure [317]S6, Supporting Information).
   Pathway enrichment analyses further demonstrated consistent alterations
   in cellular components, biological processes, and molecular functions
   across models. Notably, the [318]GSE118633 dataset corroborated
   pathway‐level changes, reinforcing the validity of our findings. These
   results not only support the robustness of our Tet‐On‐inducible
   progerin‐expressing HeLa cell model in recapitulating key molecular
   features of HGPS but also suggest that progerin expression
   differentially regulates gene expression in a substrate
   stiffness‐dependent manner by modulating multiple signaling pathways
   (Figure [319]4D–H; Figure [320]S6, Supporting Information). In
   addition, ATAC‐seq revealed that the progerin‐induced alteration of
   transcription factors is also responsive to substrate stiffness
   (Figure [321]6D–R), further suggesting that progerin‐induced nuclear
   wrinkling amplifies chromatin dynamics by promoting heterochromatin
   clustering to regulate chromatin accessibility.
   Intact LMNA‐producing normal cells typically transmit cytoskeletal
   tension to the nuclear membrane via the stable expression of LINC
   complex‐associated molecular components.^[ [322]^16 ^] Therefore, a
   strong correlation between cytoskeletal tension and the force applied
   to the nucleus has been detected.^[ [323]^59 ^] However, we noted a
   mismatch between the reduced nuclear tension and enhanced cytoskeletal
   tension in doxycycline‐induced progerin‐expressing cells (Figure [324]2
   vs Figure [325]8). This could be due to the formation of the LMNA
   mutant‐progerin, which could disrupt the LMNA‐mediated molecular
   connection. Consequently, the cytoskeletal forces may not be properly
   transmitted into the nucleus, and vice versa. The mechanical imbalance
   between outer nuclear cytoskeletal tension and LINC‐mediated nuclear
   tension in progerin‐expressing cells further highlights the critical
   role of LMNA‐dependent molecular connections between the NE and
   cytoskeleton. Notably, we showed that doxycycline‐induced progerin
   expression upregulated LINC complex‐associated genes (Figure [326]7).
   Previous studies have shown that progerin predominantly associates with
   SUN1, but not with SUN2, in LMNA^−/− mouse embryonic fibroblasts,^[
   [327]^28 ^] and that the stable expression of progerin proportionally
   increases the level of SUN1, but not SUN2, in progerin‐expressing
   NIH3T3 fibroblasts.^[ [328]^26b ^] In an in situ PLA assay, we observed
   that individual interactions between LMNA and SUN1 were increased in
   response to progerin expression, while the molecular interaction
   between SUN1 and its binding partner nesprin 2 remained unchanged
   (Figure [329]7O–T). Combining these results, we propose a new model
   that fills the missing link between myosin activity‐dependent
   cytoskeletal tension and SUN1‐mediated nuclear tension (Figure [330]8).
   Our findings suggest that progerin‐induced nuclear deformation is
   mediated by reduced nuclear tension, accompanied by SUN1
   upregulation‐dependent myosin tension, rather than by F‐actin
   connectivity through SUN1–nesprin 2 bridging. Validation in both
   SUN1‐depleted Tet‐On HeLa cells and HGPS patient‐derived fibroblasts
   established that our HeLa cell models are consistent with those from
   primary HGPS fibroblasts. This concordance reinforces that
   SUN1‐dependent nuclear tension plays a critical role in regulating
   progerin‐induced nuclear deformation, highlighting the mechanistic
   relevance of our cell model to recapitulate key aspects of pathological
   features of HGPS (Figure [331]9).
   Taken together, our results suggest that i) progerin expression,
   reduced NE tension, and nuclear wrinkling are mechanosensitive to
   changes in substrate stiffness, ii) progerin expression disrupts the
   LMNA‐mediated force balance between cytoskeletal force and nuclear
   tension, iii) the SUN1–LMNA interaction mediates force transmission
   from the cytoskeleton to the nuclear interior, and iv) reduced nuclear
   tension enhances chromatin dynamics, thereby regulating
   progerin‐induced mechanosensitive signaling pathways. The key findings
   regarding the SUN1‐dependent nuclear tension and chromatin remodeling
   have important translational implications, as they suggest that
   manipulation of nuclear tension could provide a novel therapeutic
   strategy to mitigate the effects of progerin‐induced nuclear
   deformation. These results could pave the way for developing targeted
   interventions that modulate nuclear mechanics to treat progeria as well
   as diseases associated with defects in nuclear architecture.
4. Experimental Section
Cell Culture and Drug Treatment
   HeLa cells (purchased from Korean Cell Line Bank, Seoul, Republic of
   Korea) were cultured in T25 rectangular canted neck cell culture flasks
   (Falcon, 353108) containing 2 mL of Dulbecco's Modified Eagle's Medium
   (DMEM, Corning, 10‐013‐CV) supplemented with 10% fetal bovine serum
   (FBS, Merck, TMS‐031‐BKR) and 1% penicillin–streptomycin (Thermo,
   15140122) at 37 °C with 5% CO[2] in a humidified incubator. Human
   dermal fibroblasts obtained from a 3‐year‐old normal healthy individual
   (GM05565) and a 3‐year‐old HGPS patient (AG06917) were purchased from
   Coriell Cell Repositories (Camden, NJ). Fibroblast cells were cultured
   in DMEM supplemented with 15% FBS and 1% penicillin–streptomycin at 37
   °C with 5% CO[2] in a humidified incubator. Culture media used in this
   work were refreshed every 2–3 days. To inhibit myosin II activity, HeLa
   cells were treated with 10 and 15 µm blebbistatin (Sigma‐Aldrich,
   B0560) for 1 h. Control cells, without blebbistatin, were treated with
   DMSO (Merck, 317275) and incubated under identical conditions. After
   DMSO or blebbistatin treatment, fresh culture medium was added prior to
   immunostaining and imaging.
Plasmids and Subcloning
   To generate stable cell lines with doxycycline‐inducible expression of
   wild type LMNA or Δ50 LMNA tagged with/without mCherry, PCR‐amplified
   DNA sequences for mCherry‐LMNA (Addgene, #55068), mCherry‐Δ50 LMNA
   (modified from Addgene, #17653), and Δ50 LMNA (modified from Addgene,
   #17653) were inserted into PiggyBac XLone‐GFP plasmid (Addgene, #96930)
   after digesting with KpnI, SpeI to replace the GFP sequence.^[ [332]^36
   ^] To deliver genetic cargo into the genome, blasticidin (Thermo
   Fisher, [333]R21001)‐resistant gene‐containing PiggyBac transposon was
   used.^[ [334]^60 ^] Telomeric repeat‐binding factor 2 (TRF2)‐GFP
   plasmid used for analyzing chromatin dynamics was modified.
   PCR‐amplified DNA sequences for TRF2 were obtained using TRF2‐IRES‐eGFP
   (Addgene, #19798), and PCR‐amplified DNA sequences for pcDNA3‐EGFP
   (Addgene, #13031). The TRF2 and pcDNA3‐EGFP were ligated using the
   NEBuilder HiFi DNA assembly master mix (NEB, E2621L) to construct the
   EGFP‐TRF2 plasmid.
Tet‐On System for Doxycycline‐Inducible Gene Expression
   HeLa cells were seeded in T25 rectangular canted neck cell culture
   flasks to reach 70–80% confluency after culturing for 24 h in DMEM
   supplemented with 10% FBS and 1% penicillin–streptomycin at 37 °C and
   5% CO[2] in a humidified incubator. The cells were then transfected
   with the generated plasmids (1 µg µL^−1; mCherry‐LMNA, mCherry‐Δ50
   LMNA, or Δ50 LMNA) and the PiggyBac transposase (0.5 µg µL^−1; System
   Biosciences, PB210PA‐1) using Lipofectamine 3000 (Thermo Fisher,
   L3000015) following the supplier's instructions. After 72 h of
   incubation, the cells were replated onto 96‐well cell culture plates
   (SPL, 30096) and selected with 5 µg mL^−1 blasticidin for 7 d to
   generate stable cell lines expressing the respective target genes. To
   induce gene expression, HeLa cells transfected with mCherry‐lamin A‐,
   mCherry‐Δ50 lamin A‐, or Δ50 lamin A were treated with 2 µg mL^−1
   doxycycline (Merck, D9891).
Immunofluorescence and Time‐Lapse Live Cell Monitoring
   After growing for 24 h in a glass bottom dish (SPL, 101350) coated with
   0.2 mg mL^−1 type‐I rat tail collagen (Corning, 354236) diluted in 0.2
   N acetic acid, the cells were fixed with 4% paraformaldehyde
   (Biosesang, PC2031‐100‐00) for 10 min at 4 °C, permeabilized with 0.1%
   Triton X‐100 (Merck, T8787) for 10 min, and then blocked with
   phosphate‐buffered saline (PBS, Corning, 21‐031‐CV) supplemented with
   10% FBS for 30 min at room temperature (RT). Fixed cells were incubated
   with primary antibodies for 1 h at RT. After washing with PBS three
   times, secondary antibodies with 4′,6‐diamidino‐2‐phenylindole
   dihydrochloride (DAPI, Sigma‐Aldrich, D1306) and Alexa Fluor 488
   phalloidin (Thermo, A12379) were added. Primary antibodies used in this
   study are as follows: anti‐Progerin (1:200, Abcam, ab66587), anti‐Lamin
   B1 (1:500, Abcam, ab16048), anti‐SUN1 (1:200, Millipore, ABT273),
   anti‐Nesprin 2 (1:100, Sigma‐Aldrich, MABC86), anti‐phospho‐myosin
   light chain 2 (Ser19) (1:200, CST, 3675), and di/tri‐methyl‐histone H3
   (Lys9) (1:100, CST, 5327). Secondary antibodies used are as follows:
   Goat anti‐Mouse IgG Heavy and Light Chain Antibody DyLight 488
   Conjugated (Bethyl, A90‐116D2), Sheep anti‐Rabbit IgG Heavy and Light
   Chain Antibody DyLight 488 Conjugated (Bethyl, A120‐100D2), Goat
   anti‐Mouse IgG‐Heavy and Light chain Antibody DyLight 594 Conjugated
   (Bethyl, A90‐116D4), Sheep anti‐Rabbit IgG‐Heavy and Light chain
   Antibody DyLight 650 Conjugated (Bethyl, A120‐100D5). All samples were
   imaged using confocal laser microscopy (A1R, Nikon) through 20x plan
   lens with a z‐stack of 0.4 µm or 60× oil lens with a z‐tack of 0.2 µm
   or using fluorescence microscopy (Ti2, Nikon). For time‐lapse live cell
   monitoring, HeLa cells expressing mCherry‐Δ50 lamin A were seeded onto
   a glass bottom dish and imaged through a 20× plan lens using a confocal
   microscope equipped with a stage‐top incubator (Okolab, Italy). Live
   cell images were captured every 20 min for 36 h to monitor progerin
   expression and NE wrinkling, and every 30 min for 16 h to collect
   z‐stacked NE wrinkling images. Z‐stacked time‐lapse confocal images
   were reconstructed and 3D images were rendered using NIS elements
   software (Nikon).
Nuclear Morphometry and Protein Content Measurement
   High‐throughput cell phenotyping analysis was employed to assess cell
   and nuclear size, nuclear aspect ratio, nuclear circularity, and
   protein expression at a single‐cell level.^[ [335]^61 ^] Briefly,
   immunostained cells were autofocused through the DAPI channel, and
   multiple images were automatically captured on a scale of 7 × 7 in two
   or three channels with DAPI, fluorescein‐5‐isothiocyanate (FITC), and
   tetramethylrhodamine through a 20× plan lens using a fluorescence
   microscope. Image analysis was conducted using a customized program
   coded in MATLAB (MathWorks Laboratory). In the acquired images, a
   threshold value was applied to each FITC and DAPI channel image to
   separate the cell and nuclear regions, prior to calculation of the
   fluorescence intensity. To determine the nuclear deformation, the
   fractional occupancy of nuclear wrinkles shown as the bright regions in
   the nucleus marked by expression of mCherry‐Δ50 lamin A or
   immunostaining of Lamin B1 was quantified. Specifically, the degree of
   NE wrinkling was calculated using the following equation
   [MATH: NEwrinkling%=NuclearwrinklingareaNuclearspreadingarea×100 :MATH]
   (1)
   where the nuclear wrinkling area and the nuclear spreading area
   indicate the nuclear region showing a higher intensity value than the
   average intensity value of the entire nucleus and the area of the
   fluorescence‐marked entire nucleus, respectively.
Preparation of Polyacrylamide Hydrogel Substrates
   The surface of glass bottom dishes was pretreated with (3‐aminopropyl)
   trimethoxysilane (Sigma‐Aldrich, 281778) for 5 min and then with 0.5%
   glutaraldehyde (Sigma‐Aldrich, G6257) for 30 min. After washing with
   deionized water, acrylamide/bis‐acrylamide solution containing ammonium
   persulfate (Sigma‐Aldrich, A3678) and tetramethylenediamine
   (Invitrogen, 15524‐010) was added onto the surface‐modified glass
   bottom dish. Dichlorodimethylsilane (Sigma‐Aldrich, 40140) precoated
   cover slips were immediately placed onto the droplets to form a flat
   PAG. PAGs were activated with Sulfo‐SANPAH (Thermo Fisher, 22589),
   followed by coating with type‐I rat tail collagen overnight at 4 °C. To
   modulate the stiffness of the PAG, the concentration ratios of
   acrylamide and bis‐acrylamide were varied as follows: 5% acrylamide +
   0.06% bis‐acrylamide (E ≈ 1.37 kPa) and 10% acrylamide + 0.3%
   bis‐acrylamide (E ≈ 34 kPa). The elastic moduli of the gels were
   adapted from previous studies.^[ [336]^62 ^]
Construction of Nuclear Tension Sensor
   To measure nuclear tension, a previously developed nesprin tension
   sensor was modified (Addgene, #68127).^[ [337]^63 ^] Briefly, the
   FRET‐based nesprin tension sensor consisted of mTEP1 and venus, which
   were separated by a 40‐amino acid elastic linker flanked by XhoI and
   NotI restriction sites. To modify the nesprin tension sensor to
   accommodate the confocal microscopy wavelength, the mTEP1 and venus
   were replaced with enhanced green fluorescence protein (EGFP) and
   DsRed, respectively, using the pQCXI Puro DsRed‐LC3‐GFP plasmid
   (Addgene, #31182).
FRET Imaging and Analysis
   To determine nuclear tension, cells cultured on glass bottom dishes
   were transfected with a modified nesprin tension sensor construct. EGFP
   and DsRed were excited by 488 and 561 nm lasers, respectively. FRET
   signals of cells expressing the nesprin tension sensor were captured at
   488 nm excitation wavelength through a 60× oil lens using the Nikon A1R
   confocal laser microscope equipped with a stage‐top incubator. Images
   were acquired on the same day at a fixed gain and laser intensities in
   each channel, and then analyzed using NIS‐Element software. The FRET
   ratio was defined as the ratio of energy transferred from the donor to
   the acceptor depending on the distance between the donor and acceptor
   proteins, as described by the following equation
   [MATH: FRETratio=IntensityofDsRed−Backgr
   oundofDsRedIntensityofEGFP−Backgro
   undofEGFP
   mrow> :MATH]
   (2)
   The FRET ratio was color‐coded for better visibility.
Computational Analysis of Nuclear Deformation
   A mechanical model of a soft spherical elastic thin shell was developed
   to characterize the deformation of nuclear surface. The spherical shell
   is discretized into many (5120 in this work) identical triangle face
   elements. Each face element contains 3 vertices and 3 sides, for the
   entire sphere shell, excluding repeated counted vertices, there exist a
   total of 2562 independent vertices. According to Euler's formula of
   polyhedral
   [MATH: V−E+F=2 :MATH]
   (3)
   5120 independent sides are defined, where V is the number of vertices,
   E is the number of sides, and F is the number of faces.
   The deformation and motion of the shell are described by the vertices,
   and the motion of each vertex follows the over damped Langevin equation
   [MATH: ηdridt=−∇i<
   mi>U+Fir+Pie :MATH]
   (4)
   where η is the damping coefficient, r [i] is the position vector of
   vertex i. U is the total shape potential energy of the entire shell,
   and ∇ [i] means to derive the gradient in terms of r [i] . F [i] ^r is
   the repulsive force exerting onto vertex i by other vertices that are
   too close to it and P [i] ^e is the external pressure exerting onto
   vertex i from the environment.
   Using the expression of charged elastic shell,^[ [338]^64 ^] the shape
   potential energy U of the sphere shell is written as
   [MATH: U=Ue+Ub+UV :MATH]
   (5)
   with
   [MATH: Ue=∑i=1EKe2li−l<
   /mi>02
   mrow> :MATH]
   (6)
   [MATH: Ub=∑i=1EKb2n^i,1−n^i,22
   mrow> :MATH]
   (7)
   [MATH: UV=∑i=1FKV2Vi−V<
   /mi>02
   mrow> :MATH]
   (8)
   where U [e] represents the stretching energy, generated from the
   extension of shortening of all sides, K [e] is the stretching modulus,
   l[i] is the actual length of side i, l [0] is the initial side length
   shared by all sides. U [b] represents the bending energy, generated
   from the bending of the membrane featured by the dihedral angle between
   each two adjacent faces, K [b] is the bending modulus,
   [MATH: n^i,1 :MATH]
   and
   [MATH: n^i,2 :MATH]
   are the unit outward normal vectors of the 2 adjacent faces of side i.
   U [e] and U [b] describe the mechanical properties of the membrane
   itself. U [V] represents the volume energy, generated from the
   elasticity of the contents of the shell, K [V] is the volume modulus,
   V[i] is volume element i’s volume defined by face i, and V [0] is the
   initial volume shared by all elements.
   F [i] ^r represents the repulsive forces between 2 very close vertices,
   and the traditional 6–12 law is adopted to characterize the molecular
   level interactions
   [MATH: Fir=−Kr∑jε
   Ci1ri−r<
   /mi>j6−1ri−r<
   /mi>j12
   ri−r<
   /mi>j :MATH]
   (9)
   where K [r] is the repulsion coefficient, C[i] represents the set of
   other vertices whose distance from vertex i is less than a threshold
   distance D.
   The cytoskeleton in connected to the nuclear lamina through the LINC
   complexes, and the LINC complexes are regarded as a molecular spring.
   Once the progerin is expressed, the distance between cytoskeleton and
   the nuclear lamina gets reduced (refer to FRET Imaging and Analysis),
   leading to a reduced nuclear tension, thus the spring can generate
   greater compressive pressure between the cytoskeleton and nucleus
   membrane. Hence, a uniform and constant external pressure term P [i] ^e
   is used to model the force exerting onto the shell (nucleus membrane)
   by the environment (cytoskeleton)
   [MATH: Pie=−Pie∇i<
   mi>V :MATH]
   (10)
   where P[i] ^e is the scalar value of P [i] ^e, and V is the total
   volume of the shell. With increasing substrate stiffness, the pressure
   also increases, therefore, P[i] ^e is set higher on stiffer substrate.
   Specifically, P[i] ^e = 500P [0], 600 P [0], and 750 P [0] are set on
   soft, medium, and stiff substrates, respectively. All parameters and
   values used in the model were listed in Table [339]1 .
Table 1.
   Parameters for mechanical model.
   Parameter Numerical value Refs.
   F 5120 [[340]71]
   V 2562
   E 7680
   K [e] 1000 [[341]64]
   K [b] 4
   K [V] 3
   K [r] 1 × 10^−13
   [MATH: Pie :MATH]
   500 on soft substrate [[342]72]
   600 on medium substrate
   750 on stiff substrate
   η 1 [[343]73]
   dt 0.005
   D 0.9
   [344]Open in a new tab
RNA Sequencing
   Cells cultured on glass and 1.37 kPa PA gel were used to extract total
   RNA using the RNeasy Mini Kit (QIAGEN, 74104). A library was
   independently prepared with 1 µg of total RNA for each sample by
   Illumina TruSeq Stranded mRNA Sample Prep Kit (Illumina, Inc., San
   Diego, CA, USA, #20020595). Libraries were quantified using KAPA
   Library Quantification kits for Illumina Sequencing platforms according
   to the qPCR Quantification Protocol Guide (KAPA BIOSYSTEMS, #KK4854)
   and qualified using a TapeStation D1000 ScreenTape (Agilent
   Technologies, # 5067–5582). Indexed libraries were then subjected to
   Illumina NovaSeq (Illumina, Inc., San Diego, CA, USA), and paired‐end
   (2 × 100 bp) sequencing was performed by Macrogen Inc. (South Korea).
   Reference genome sequences and gene annotation data were downloaded
   from the NCBI Genome Assembly and RefSeq databases, respectively. The
   aligned data (SAM file format) were sorted and indexed using SAM tools
   v 1.9. After alignment, the transcripts were assembled and quantified
   using StringTie v2.1.3b.^[ [345]^65 ^] Gene‐ and transcript‐level
   quantification were calculated as the raw read count, Fragments Per
   Kilobase of transcript per million mapped reads, and transcript per
   million mapped reads. The statistical significance of the DEGs was
   determined using the edgeR most exact, and the fold change and p‐value
   were extracted from the most exact results. All p‐values were adjusted
   using the Benjamini–Hochberg algorithm to control the false discovery
   rate. Hierarchical clustering of the log‐transformed values for
   significant genes was performed using these parameters (distance metric
   = Euclidean distance; linkage method = complete). Gene‐enrichment and
   functional annotation analysis for significant genes were carried out
   using gProfiler (Raudvere, Uku, et al. 2019,
   [346]https://biit.cs.ut.ee/gprofiler/orth) against Gene Ontology (GO)
   database. Adjusted p‐values reported from the gProfiler results were
   derived using a one‐sided hypergeometric test and corrected using the
   Benjamini–Hochberg method.
Assay for Transposase‐Accessible Chromatin Using Sequencing (ATAC Sequencing)
   For ATAC sequencing, samples were prepared according to previously
   established methods. Briefly, cells were detached from the glass and
   1.37 kPa polyacrylamide hydrogel; 100 000 cells were prepared. The
   cells were then lysed in a cold lysis buffer. The nuclei concentration
   was determined using LUNA‐FL Automated Fluorescence Cell Counter (logos
   biosystems) and the nuclei morphology was examined using microscopy.
   Immediately after lysis, the transposition reaction was continued using
   Tagment DNA TDE1 Enzyme and Buffer Kit (Illumina). Nuclei (50 000
   cells) were resuspended in the transposition reaction mixture by
   incubating for 30 min at 37 °C. Immediately after transposition, cells
   were purified using the Qiagen MinElute PCR Purification Kit. ATAC‐seq
   was performed by Macrogen, Inc. (South Korea). The cleaned reads were
   aligned to the human genome (GRCh38) using Bowtie2.^[ [347]^66 ^] The
   mapped data (in SAM file format) were sorted and indexed using SAM
   tools (version 1.9).^[ [348]^67 ^] After removing the reads aligned to
   the mitochondrial genome from the indexed BAM file, duplicate reads
   were removed using MarkDuplicates in Picard (version 0.118). Peaks in
   the aligned sequence data were identified using a model‐based analysis
   of ATAC‐seq (MACS2 version 2.1.1.20160309).^[ [349]^68 ^] The algorithm
   empirically models the length of ATAC‐seq fragments from sequence data,
   considering local genomic biases in the distribution of mapped reads.
   The following parameters were used: macs2 callpeak‐t ATAC‐seq. bam‐g
   hs–bdg–nolambda–keep‐dup all–broad. Among the called peaks, those that
   overlapped with the ENCODE blacklisted regions were excluded.
   ChIPseeker (version 1.22.1),^[ [350]^69 ^] a Bioconductor package in R
   (version 4.2.2) used to facilitate batch annotation of enriched peaks,
   was used to identify nearby genes and transcripts from the peaks
   obtained by MACS2. The HOMER software was used to discover de novo TF
   binding motifs within ATAC‐seq‐defined regions. An integrative genome
   viewer (IGV; [351]http://igv.org/) was used to visualize the genome
   track corresponding to the gene locus.
Chromatin Dynamics Analysis
   Chromatin dynamics were analyzed by tracking telomere movement in
   three‐year normal healthy (GM00565) and HGPS fibroblast cell lines
   (AG06917) using the EGFP‐TRF2 plasmid. Briefly, 2.0 µg of EGFP‐TRF2 in
   Opti‐MEM (Thermo Fisher, 31985062) was mixed with 2.0 µL of
   Lipofectamine 3000 and 4.0 µL of P3000 transfection reagent in
   Opti‐MEM. The mixture was gently incubated for 20 min at RT, followed
   by a 4 h incubation at 37 °C and 5% CO[2] in a humidified incubator.
   Next, transfected EGFP‐TRF2 cells were imaged at 0.07 s intervals for 5
   min. Trajectory images were analyzed using the LIM Tracker Plugin in
   the Plugins menu of ImageJ, and MSD analysis was performed using a
   custom MATLAB script.
Immunoblotting
   Cellular lysates were prepared using RIPA buffer (Thermo Fisher
   Scientific, 89901) supplemented with a 1% protease inhibitor cocktail
   (Sigma, P8340). These preparations were then electrophoresed using
   SDS‐PAGE, and the resolved proteins were transferred to a
   polyvinylidene difluoride membrane (Thermo Fisher, LC2002). Nonspecific
   interactions were blocked using a protein‐based blocking reagent
   (Invitrogen, T2015) for 1 h. The membranes were then incubated with
   primary antibodies against H3K9me2/3 (CST, 5327), SUN1 (Proteintech,
   24568), Nesprin 2 (Abcam, ab233034), and GAPDH (Sigma, G8795) overnight
   at 4 °C. Subsequently, membranes were incubated with the appropriate
   secondary antibodies, goat anti‐mouse IgG heavy and light chain
   antibody HRP Conjugated (Bethyl, A90‐116P), and goat anti‐rabbit IgG
   heavy and light chain antibody HRP Conjugated (Bethyl, A120‐101P) at RT
   for 1 h. Proteins were visualized using an ECL kit (Thermo Fisher
   Scientific, 32106).
Proximity Ligation Assay (PLA) Analysis
   Subconfluent cells were fixed on the glass bottom dish with 4%
   paraformaldehyde for 10 min at 4 °C and permeabilized with 0.1% Triton
   X‐100 for 10 min at RT. After blocking with Doulink Blocking Solution
   (Sigma‐Aldrich, DUO82007) for 60 min at 37 °C in a humidity chamber,
   samples were incubated overnight at 4 °C with anti‐SUN1 (1:200,
   Millipore, ABT273), anti‐Lamin A/C (1:200, Cell Signaling, 4777S), or
   anti‐Nesprin 2 (1:100, Sigma‐Aldrich, MABC86). Cells were washed with
   1× wash Buffer A (Sigma‐Aldrich, DUO82049) and incubated in preheated
   humidity chamber for 1 h at 37 °C with anti‐PLUS and anti‐MINUS PLA
   probes diluted 1:5 in the Duolink Antibody Diluent (Sigma‐Aldrich,
   DUO92013). Samples were then incubated in a humidity chamber at 37 °C
   with ligation solution diluted 1:40 in the 1× ligation buffer. After 30
   min, ligation buffer diluted 1:80 in 1× amplification buffer was added
   to the samples, followed by incubation for 100 min at 37 °C in
   preheated humidity chamber. Samples were mounted using Duolink in situ
   mounting medium (Sigma‐Aldrich, DUO82040) with DAPI. The cells were
   imaged using confocal laser microscopy through a 60× oil lens. To
   confirm the level of protein–protein interaction, the number and total
   intensity of red dots per cell were assessed based on the fluorescence
   signal.^[ [352]^70 ^]
siRNA‐Mediated Knockdown
   To knockdown SUN1, siRNA targeting SUN1 (Santa Cruz Biotechnology,
   sc‐106672) and control siRNA (Santa Cruz Biotechnology, sc‐37007) were
   delivered to progerin‐expressing HeLa cells or HGPS fibroblast cells
   (Coriell Cell Repositories), according to the manufacturer's
   instructions. Briefly, 75 pmol siRNA in Opti‐MEM (Thermo Fisher,
   31985062) and 7.5 µL Lipofectamine 3000 transfection reagent in
   Opti‐MEM were gently mixed and incubated for 5 min at RT.
   Progerin‐expressing HeLa cells at 70–80% confluency on glass bottom
   dishes were transiently transfected with siControl or siRNA targeting
   SUN1 for 24 h. After removing the transfection mixture and replacing it
   with fresh growth medium, cells were immunostained for analysis.
Data Processing and Statistical Analysis
   All statistical analyses were performed using GraphPad Prism (GraphPad
   software, USA), and statistical significance was analyzed using
   unpaired t‐test or Student's t‐test when two groups were compared.
   One‐way analysis of variance (ANOVA) using Tukey's test or Bonferroni's
   post hoc test was used for pairwise comparisons between multiple
   groups. Error bars represent the standard deviation (S.D.) or standard
   error of the mean (S.E.M.), as indicated. Significance levels are
   indicated with asterisks in each figure as follows: *: p < 0.01; **: p
   < 0.005; ***: p < 0.001; ****: p < 0.0001.
Conflict of Interest
   The authors declare no conflict of interest.
Author Contributions
   J.P., J.J., and K.X. contributed equally to this work. J.P. and J.J.
   performed and analyzed all experiments unless otherwise specified and
   cowrote the manuscript. K.X. developed computational model to simulate
   the nuclear wrinkling. S.L. and S.H. analyzed the protein contents and
   chromatin dynamics in the human dermal fibroblasts. J.J., Y.L., and
   W.C. developed the Tet‐On system. W.C., B.L., and S.H.K. provided a
   conceptual design of the experimental settings. K.X. and B.L. cowrote
   the manuscript. D.K. supervised the project, designed the experiments,
   and wrote the manuscript.
Supporting information
   Supporting Information
   [353]ADVS-12-2502375-s002.docx^ (8.4MB, docx)
   Supplemental Movie 1
   [354]Download video file^ (499KB, avi)
   Supplemental Movie 2
   [355]Download video file^ (610.9KB, avi)
   Supplemental Movie 3
   [356]Download video file^ (1.1MB, avi)
   Supplemental Movie 4
   [357]Download video file^ (1.7MB, avi)
Acknowledgements