COMMENTARY - Journal of Cell Science · 2014. 7. 8. · the concentration of ECM components and...

11
Journal of Cell Science COMMENTARY The nuclear lamina is mechano-responsive to ECM elasticity in mature tissue Joe Swift and Dennis E. Discher* ABSTRACT How cells respond to physical cues in order to meet and withstand the physical demands of their immediate surroundings has been of great interest for many years, with current research efforts focused on mechanisms that transduce signals into gene expression. Pathways that mechano-regulate the entry of transcription factors into the cell nucleus are emerging, and our most recent studies show that the mechanical properties of the nucleus itself are actively controlled in response to the elasticity of the extracellular matrix (ECM) in both mature and developing tissue. In this Commentary, we review the mechano-responsive properties of nuclei as determined by the intermediate filament lamin proteins that line the inside of the nuclear envelope and that also impact upon transcription factor entry and broader epigenetic mechanisms. We summarize the signaling pathways that regulate lamin levels and cell-fate decisions in response to a combination of ECM mechanics and molecular cues. We will also discuss recent work that highlights the importance of nuclear mechanics in niche anchorage and cell motility during development, hematopoietic differentiation and cancer metastasis, as well as emphasizing a role for nuclear mechanics in protecting chromatin from stress-induced damage. KEY WORDS: Cell mechanics, Mechanotransduction, Extra-cellular matrix, Nucleus, Nucleoskeleton, Proteostasis Introduction Evolution has likely driven our tissues and organs to fulfill their roles with optimal efficiency and, of course, viability. Mature tissues need to be particularly resistant to the mechanical demands of an active life. Our bones, cartilage, skeletal muscle and heart tissues are stiff, making them robust and resistant to routine physical exertion, such as walking or running, during which they are subjected to high-frequency shocks, stresses and strains. With every heartbeat, the left ventricular wall experiences a 20% radial strain (Aletras et al., 1999), and local strains of ,20% also occur in the cartilage of knee joints with every step (Guilak et al., 1995). Tissue-level deformations might even be amplified within cells and their nuclei (Henderson et al., 2013). Our softer tissues have less need for robustness because their function does not require them to bear load. Furthermore, some of our softest tissues, such as brain and marrow, are protected from shocks and stresses by our bones. Nonetheless, when soft tissues are subjected to impact, such as during a collision of heads in American football or rugby, occurrences of rapid straining can cause lasting damage (Viano et al., 2005). We have recently sought to characterize the composition of cells and extra-cellular matrix (ECM) in tissues of increasing stiffness, and by implication, in tissues that are subjected to the greatest stress (Swift et al., 2013b). A close correlation between the concentration of ECM components and bulk tissue elasticity was discovered for mouse tissues (Swift et al., 2013b). More surprisingly, we also discovered a systematic scaling between tissue elasticity and the concentration of lamins in the nucleoskeleton that was partially re-capitulated in cultured cell systems (Swift et al., 2013b). Corresponding changes to nuclear mechanical properties associated with this tuning of lamina composition suggest that this response might act to protect the chromatin cargo of the nucleus from shocks that are transmitted through the surroundings, across the cytoskeleton and into the nucleus. An active regulation of cell or ECM composition in response to the environment implies feedback into pathways of protein turnover and remodeling, or control of the rate of new protein production. Responsive matching of mechanical properties to physical demands has classically been described as a ‘mechanostat’ in the context of bone regulation (Frost, 1987), but a recent explosion in mechanobiological studies has uncovered a host of other mechanically sensitive cellular phenomena, including contraction (Discher et al., 2005), migration (Hadjipanayi et al., 2009b; Winer et al., 2009; Raab et al., 2012), proliferation (Hadjipanayi et al., 2009a; Klein et al., 2009; Lo et al., 2000), differentiation (Engler et al., 2004; Engler et al., 2006) and apoptosis (Wang et al., 2000). However, despite great recent progress, questions of how mechanical signals are transduced into specific transcriptional or regulatory pathways continue to challenge the field. The lamina is a network structure formed from intermediate filament lamin proteins, and it lies just inside the nuclear envelope and interacts with both chromatin and the cytoskeleton (Fig. 1A). In the somatic cells of humans, mice and most vertebrates, the main forms of lamin protein are expressed from three genes: lamin-A and lamin-C are alternatively spliced products of the LMNA gene, and they are collectively referred to as ‘A-type’ lamins; lamin-B1 and lamin-B2 are encoded by the LMNB1 and LMNB2 genes, respectively, and are known as ‘B- type’ lamins. The lamins share structural features and, indeed, have some similarity in amino acid sequence, but they differ in their post-translational modification, with B-type lamins permanently appended by a farnesyl group that is cleaved from mature lamin-A (reviewed, for example, by Dechat et al., 2010; Ho and Lammerding, 2012). Like other intermediate filament proteins, such as keratin and vimentin, the lamins form coiled- coil parallel dimers that assemble into higher-order filamentous structures that fulfill important structural roles (Herrmann et al., 2009). Molecular and Cell Biophysics Lab, University of Pennsylvania, Philadelphia, PA 19104, USA. *Author for correspondence ([email protected]) ß 2014. Published by The Company of Biologists Ltd | Journal of Cell Science (2014) 127, 3005–3015 doi:10.1242/jcs.149203 3005

Transcript of COMMENTARY - Journal of Cell Science · 2014. 7. 8. · the concentration of ECM components and...

  • Jour

    nal o

    f Cel

    l Sci

    ence

    COMMENTARY

    The nuclear lamina is mechano-responsive to ECM elasticity inmature tissue

    Joe Swift and Dennis E. Discher*

    ABSTRACT

    How cells respond to physical cues in order to meet and withstand

    the physical demands of their immediate surroundings has been of

    great interest for many years, with current research efforts focused

    on mechanisms that transduce signals into gene expression.

    Pathways that mechano-regulate the entry of transcription factors

    into the cell nucleus are emerging, and our most recent studies

    show that the mechanical properties of the nucleus itself are actively

    controlled in response to the elasticity of the extracellular matrix

    (ECM) in both mature and developing tissue. In this Commentary,

    we review the mechano-responsive properties of nuclei as

    determined by the intermediate filament lamin proteins that line

    the inside of the nuclear envelope and that also impact upon

    transcription factor entry and broader epigenetic mechanisms.

    We summarize the signaling pathways that regulate lamin levels

    and cell-fate decisions in response to a combination of ECM

    mechanics and molecular cues. We will also discuss recent work

    that highlights the importance of nuclear mechanics in niche

    anchorage and cell motility during development, hematopoietic

    differentiation and cancer metastasis, as well as emphasizing a role

    for nuclear mechanics in protecting chromatin from stress-induced

    damage.

    KEY WORDS: Cell mechanics, Mechanotransduction, Extra-cellular

    matrix, Nucleus, Nucleoskeleton, Proteostasis

    IntroductionEvolution has likely driven our tissues and organs to fulfill their

    roles with optimal efficiency and, of course, viability. Mature

    tissues need to be particularly resistant to the mechanical

    demands of an active life. Our bones, cartilage, skeletal muscleand heart tissues are stiff, making them robust and resistant to

    routine physical exertion, such as walking or running, during

    which they are subjected to high-frequency shocks, stresses and

    strains. With every heartbeat, the left ventricular wall experiencesa 20% radial strain (Aletras et al., 1999), and local strains of

    ,20% also occur in the cartilage of knee joints with every step(Guilak et al., 1995). Tissue-level deformations might even be

    amplified within cells and their nuclei (Henderson et al., 2013).Our softer tissues have less need for robustness because their

    function does not require them to bear load. Furthermore, some of

    our softest tissues, such as brain and marrow, are protected fromshocks and stresses by our bones. Nonetheless, when soft tissues

    are subjected to impact, such as during a collision of heads in

    American football or rugby, occurrences of rapid straining cancause lasting damage (Viano et al., 2005).

    We have recently sought to characterize the composition ofcells and extra-cellular matrix (ECM) in tissues of increasingstiffness, and by implication, in tissues that are subjected to thegreatest stress (Swift et al., 2013b). A close correlation between

    the concentration of ECM components and bulk tissue elasticitywas discovered for mouse tissues (Swift et al., 2013b). Moresurprisingly, we also discovered a systematic scaling between

    tissue elasticity and the concentration of lamins in thenucleoskeleton that was partially re-capitulated in cultured cellsystems (Swift et al., 2013b). Corresponding changes to nuclear

    mechanical properties associated with this tuning of laminacomposition suggest that this response might act to protect thechromatin cargo of the nucleus from shocks that are transmitted

    through the surroundings, across the cytoskeleton and into thenucleus. An active regulation of cell or ECM composition inresponse to the environment implies feedback into pathways ofprotein turnover and remodeling, or control of the rate of new

    protein production. Responsive matching of mechanical propertiesto physical demands has classically been described as a‘mechanostat’ in the context of bone regulation (Frost, 1987), but

    a recent explosion in mechanobiological studies has uncovered ahost of other mechanically sensitive cellular phenomena, includingcontraction (Discher et al., 2005), migration (Hadjipanayi et al.,

    2009b; Winer et al., 2009; Raab et al., 2012), proliferation(Hadjipanayi et al., 2009a; Klein et al., 2009; Lo et al., 2000),differentiation (Engler et al., 2004; Engler et al., 2006) and

    apoptosis (Wang et al., 2000). However, despite great recentprogress, questions of how mechanical signals are transduced intospecific transcriptional or regulatory pathways continue tochallenge the field.

    The lamina is a network structure formed from intermediatefilament lamin proteins, and it lies just inside the nuclear

    envelope and interacts with both chromatin and the cytoskeleton(Fig. 1A). In the somatic cells of humans, mice and mostvertebrates, the main forms of lamin protein are expressed from

    three genes: lamin-A and lamin-C are alternatively splicedproducts of the LMNA gene, and they are collectively referredto as ‘A-type’ lamins; lamin-B1 and lamin-B2 are encoded by the

    LMNB1 and LMNB2 genes, respectively, and are known as ‘B-type’ lamins. The lamins share structural features and, indeed,have some similarity in amino acid sequence, but they differ intheir post-translational modification, with B-type lamins

    permanently appended by a farnesyl group that is cleaved frommature lamin-A (reviewed, for example, by Dechat et al., 2010;Ho and Lammerding, 2012). Like other intermediate filament

    proteins, such as keratin and vimentin, the lamins form coiled-coil parallel dimers that assemble into higher-order filamentousstructures that fulfill important structural roles (Herrmann et al.,

    2009).

    Molecular and Cell Biophysics Lab, University of Pennsylvania, Philadelphia,PA 19104, USA.

    *Author for correspondence ([email protected])

    � 2014. Published by The Company of Biologists Ltd | Journal of Cell Science (2014) 127, 3005–3015 doi:10.1242/jcs.149203

    3005

    mailto:[email protected]

  • Jour

    nal o

    f Cel

    l Sci

    ence

    Micropipetteaspiration

    Embryonicheart

    Log (micro-stiffness)

    Bone

    Brain

    Marrow

    Fat MuscleLo

    g (A

    -type

    :B-ty

    pe la

    min

    ) A-type lamin dominant

    B-type lamin dominant

    Lung

    Kidney

    Heart

    Liver

    Cartilage

    Log

    (type

    I co

    llage

    n)

    Log (micro-stiffness)

    High collagen stiff tissue

    Muscle

    Heart

    Lung

    Brain

    Liver

    Bone

    Cartilage

    Kidney

    Fat

    Embryonic disc

    Embryo age (days)

    Log

    (mic

    ro-s

    tiffn

    ess)

    Adult range

    Adult range

    Heart

    Brain

    Embryo age (days)

    Log

    (type

    I co

    llage

    n)

    Heart

    Brain

    Pluripotent Fibroblast

    Rel

    ativ

    e nu

    clea

    r stif

    fnes

    s

    HighA-typelamin

    LowA-typelamin

    AdultTime in culture (days)

    Nuclear envelope

    SUN proteinsNesprins

    Cytoskeletalproteins

    LINCcomplex

    ChromatinA-type laminB-type lamin

    Nuclear lamina

    Chromatin

    ChromatinNuclear poresF-actin

    Microtubules

    Nucleus

    Intermediate filaments

    Lamin-A

    Lamin-C

    Lamin proteins

    Lamin-B1

    Lamin-B2

    Ig-domain

    LMNA

    LMNB1

    LMNB2

    Coiled-coil domains

    Common region

    Very lowA-type lamin

    Stem celldifferentiation

    Very low A-type lamin

    B Lamin-A scales with the collagen-dependent stiffness of mature tissues

    C Tissue stiffens in development and so does the nucleus

    A A- and B-type lamins assemble between the nuclear envelope and chromatin

    Fig. 1. Relationship between the ECM and lamin in mature tissue and during development. (A) A-type and B-type lamins (red and blue, respectively) formjuxtaposed networks on the inside of the nuclear envelope; they are effectively located at an interface between chromatin and the cytoskeleton, the latter ofwhich is attached to the lamina through the LINC complex (shown in the magnified view). The A-type lamins, lamin-A and lamin-C, are alternativespliceoform products of the LMNA gene; the B-type lamins, lamin-B1 and lamin-B2, are protein products of LMNB1 and LMNB2, respectively (adapted fromBuxboim et al., 2010a, originally published in The Journal of Cell Science). (B) Left: the quantity of type I collagen present in tissues is proportional to tissuemicro-elasticity (Swift et al., 2013b). As collagen is one of the most prevalent proteins in the body, it is, perhaps, expected that it defines mechanical properties.Right: the ratio of A-type lamin to B-type lamin in the nuclear lamina is proportional to tissue microelasticity. The lamina contains predominantly A-type laminsin stiff tissue, whereas B-type lamins are prevalent in the nuclear lamina in soft tissue (Swift et al., 2013b). (C) Left: observations made in adult tissue areconsistent with results from the developing chick; the embryonic disc was initially soft, but divergent tissues either remained soft (e.g. brain, blue) orbecame increasingly stiff (e.g. heart, red). Inset: developing chick hearts were probed by micropipette aspiration to determine micro-elasticity. Middle: tissuestiffening during development is accompanied by increased levels of collagen and A-type lamins (Lehner et al., 1987; Majkut et al., 2013). Right: embryonic stemcells initially express negligible quantities of A-type lamin proteins, but these levels increase as the nucleus stiffens during lineage commitment (Pajerowskiet al., 2007).

    COMMENTARY Journal of Cell Science (2014) 127, 3005–3015 doi:10.1242/jcs.149203

    3006

  • Jour

    nal o

    f Cel

    l Sci

    ence

    This Commentary aims to summarize recent efforts tocharacterize the proteins that vary systematically with tissue

    stiffness. The effects of the composition of the lamina on nuclearmechanical properties will be elaborated in detail, and we willconsider the functions of the lamina in transducing mechanicalsignals from the ECM and the surroundings of the cell into

    cellular responses, both in terms of the active regulation of thelamina itself and its broader role in linking mechanotransductionpathways. Although we focus on a primarily protective function

    of nuclear lamin, there are additional regulatory consequences ofpossessing such a stiff and bulky organelle as the nucleus, and wewill summarize recent evidence that such properties limit the

    freedom of cells to move through tissue. The proximity of thelamina to heterochromatin within the nucleus has led it to bewidely associated with epigenetic regulation (e.g. Kim et al.,

    2011; Meuleman et al., 2013). This Commentary seeks tohighlight the pervasive influence of the mechanical role of thelamina and hence proposes that lamin acts as both the guardianand the gatekeeper of chromatin.

    Scaling of ECM and lamina components in mature anddeveloping tissueCollagens and other protein constituents of the ECM are the mostprevalent proteins in our bodies, largely determining themechanical properties of tissue. Collagens are found at higher

    levels in stiff mature tissues where, consistent with an expectationfor proteins to behave as ‘biological polymers’ (Gardel et al.,2004), their increased concentration is the basis of tissue elasticity

    (Fig. 1B, left panel). By using quantitative label-free massspectrometry for proteomic profiling (Swift et al., 2013a), wehave shown that collagens and other ECM-associated proteinsscale with tissue elasticity (Swift et al., 2013b). Mass

    spectrometry was also used to quantify ,100 of the mostabundant proteins in the cytoskeleton and nucleus, and we foundthat the elasticity of bulk tissue is strongly correlated with the

    composition of the nuclear lamina in terms of its content of A-type and B-type lamins (Fig. 1B, right panel). Although primarilycharacterized by the ratio between these two main families of

    lamins, the compositional scaling is dominated by a 30-foldincrease in the concentration of A-type lamin from brain to bone.Although our recent observations are broadly in agreement withan extensive literature on lamin quantification (e.g. Broers et al.,

    1997; Cance et al., 1992; Krohne et al., 1981; Röber et al., 1990),they provide a new perspective on systematic variations acrossmany tissues.

    The relationship between tissue stiffness and ECM duringdevelopment has also been determined (Majkut et al., 2013);micropipette aspiration of embryonic chick tissue showed that the

    homogeneous embryonic disk is initially very soft, withproteomic profiles indicating correspondingly low levels ofcollagen (Fig. 1C, left and middle panels). However, the

    properties of different tissues diverge during development, withthe brain remaining soft, whereas the heart stiffens as ECMproteins are deposited (Majkut et al., 2013). Cells in stiffeningtissues, such as the heart, are also likely to have correspondingly

    higher levels of lamin-A and lamin-C (Lehner et al., 1987).Nuclei in embryonic stem cells have, indeed, been shown to bevery soft and to have low levels of lamin-A and lamin-C

    (Eckersley-Maslin et al., 2013; Pajerowski et al., 2007). As thesecells commit to a lineage-specific fate, the levels of these laminsincrease, and the nucleus becomes correspondingly stiffer

    (Fig. 1C, right panel).

    Importantly, despite an apparent role in reinforcing thedecisions of animal cell fate (in conjunction with ECM

    elasticity) (Swift et al., 2013b), lamin-A and lamin-C are notessential to development, because knockout mice that lack A-typelamins still form all tissues (Sullivan et al., 1999). Likewise,lamin-B knockout mice survive embryogenesis (Kim et al.,

    2011), and embryonic stem cells can be cultured anddifferentiated without having any lamins (Kim et al., 2013).The most crucial role of lamin might therefore be to fine-tune the

    properties and regulation of maturing tissues in higher organisms,and its absence can perhaps be compensated for duringdevelopment. However, we note that the distinction here might

    be blurred; there is still a need to regulate and maintain nuclearstructure during some stages of development, such as during cellmigration, and the processes of trafficking and differentiation

    continue throughout the lifespan. Nonetheless, the fact thatlamins are not absolutely essential for cell viability is consistentwith the current notion that lamins are not expressed in yeast andplants (Dittmer and Misteli, 2011), despite the latter possessing

    genomes that are larger and more complex than those of animals.The hard cell walls of these organisms likely protect thechromatin in a manner that is not possible for animal cells with

    soft cell membranes.

    The influence of lamina composition on the mechanicalproperties of the nucleusMicropipette aspiration experiments have enabled the detailedstudy of the mechanical properties of the nucleus, which can be

    investigated by measuring the rate of deformation under pressure(Fig. 2A,B; Dahl et al., 2005; Pajerowski et al., 2007). Byexamining nuclei with different lamina compositions, it is thuspossible to approximate the characteristic contributions of A-type

    and B-type lamins to nuclear properties (Fig. 2C; Harada et al.,2014; Shin et al., 2013; Swift et al., 2013b). The nuclearmechanical response has been characterized as a combination of

    elastic (spring-like) and viscous (liquid-like, flowing) properties,with B-type lamins contributing primarily to the elastic responseand A-type lamins contributing to the viscosity (Harada et al.,

    2014; Shin et al., 2013; Swift et al., 2013b). Thus, the differencebetween a nucleus in which the lamina is stoichiometricallydominated by A-type versus B-type lamins might be akin tocomparing a balloon filled with honey to one filled with water.

    The importance of A-type lamins in maintaining nuclearstructural integrity and cell viability has been appreciated formany years (e.g. Broers et al., 2004; Lammerding et al., 2006),

    and the influence of A-type lamins on nuclear viscosity has beenmore recently demonstrated in studies where nuclei are deformedduring migration through microfluidic circuits (Rowat et al.,

    2013) or transwell pores (discussed later; Harada et al., 2014).Laminopathies are a family of diseases that are caused by

    truncations or single amino acid substitutions in A-type lamins

    (reviewed, for example, by Butin-Israeli et al., 2012; Davidsonand Lammerding, 2014; Isermann and Lammerding, 2013;Worman, 2012). These disorders include muscular dystrophies(Bonne et al., 1999), cardiomyopathies (Fatkin et al., 1999),

    lipodystrophies (Hegele et al., 2000; Shackleton et al., 2000;Speckman et al., 2000) and premature ageing (termed ‘progeria’;Merideth et al., 2008). Indeed, one of the confounding aspects of

    lamin-related disease is the mechanism by which such a widelyexpressed protein can cause tissue-specific symptoms. Althoughmuch further work is needed to resolve this question, it is broadly

    true that laminopathies cause defects in tissues where A-type

    COMMENTARY Journal of Cell Science (2014) 127, 3005–3015 doi:10.1242/jcs.149203

    3007

  • Jour

    nal o

    f Cel

    l Sci

    ence

    lamins are the dominant nuclear lamin (i.e. bone, heart, muscleand fat). However, there are exceptions – Charcot-Marie-Toothdisorder affects the nervous system (De Sandre-Giovannoli et al.,2002). Mouse models of A-type lamin knockout display defects

    in muscle and connective tissue and typically die from heartfailure several weeks after birth (Jahn et al., 2012; Kubben et al.,2011; Sullivan et al., 1999). Despite the apparently constitutive

    expression of B-type lamins, mouse models with lamin-B1 andlamin-B2 ablation progress through embryogenesis, with theireventual death being due to defects in brain development

    (Coffinier et al., 2011; Kim et al., 2011).

    Mechanisms of lamin regulationEarlier discussion has posited that lamins are closely regulated to

    match the mechanical properties of the nucleus with the physicaldemands of the tissue. In addition to being determined by theepigenetic programming required to make a given tissue or organ,

    it is also important to note that protein levels vary in response tofeedback from their surroundings. Even within bulk tissues,mechanical loading can cause inhomogeneous straining (for

    example in human articular cartilage, meniscus and ligaments;Chan and Neu, 2012), making it beneficial to have mechanisms oflamin regulation that act at the local level of individual cells.

    The many mechanisms by which the levels of lamin-A andlamin-C can be regulated are summarized in Fig. 3A. We haveshown that the transcript and protein levels of A-type lamins arehighly correlated in tissue (Swift et al., 2013b), a finding that is

    suggestive of a tight regulatory feedback. A recent study of theproteome and transcriptome in mouse fibroblasts showed thatthere are around 107 copies of A-type lamin proteins per cell –

    accounting for about 0.7% of cellular protein mass – and around200 copies of the LMNA transcript. Half-life in the cell is reportedto be ,4 days for the protein and ,20 hours for the mRNA, bothslightly higher than the cellular average for all proteins andtranscripts (Schwanhäusser et al., 2011). Measurements made onproteins in a human lung cancer cell line showed the half-life ofA-type lamin to be ,12 hours, roughly in the middle of the spanof protein half-lives recorded in the study (Eden et al., 2011).DNA methylation is an epigenetic mechanism by which gene

    activity can be regulated, but it was discounted as the foremostmeans of controlling LMNA levels – no consistent changes wereobserved in the methylation of the LMNA promoter either in arange of cell lines known to express different levels of A-type

    lamin protein (Swift et al., 2013b) or in tissues from patients withlaminopathic disorders (Cortese et al., 2007). LMNA transcriptionhas been reported to be controlled by transcription factors of the

    retinoic-acid receptor family (RAR and RXR family proteins;Okumura et al., 2004a; Okumura et al., 2004b; Olins et al., 2001;Shin et al., 2013; Swift et al., 2013b), with the resulting mRNA

    alternatively spliced to give the lamin-A and truncated lamin-Cforms. Soft tissue generally preferentially expresses the lamin-Cspliceoform (Swift et al., 2013b), and, in brain, the micro-RNAMIR-9 specifically targets and deactivates the mRNA of the

    lamin-A spliceoform (Jung et al., 2012; Jung et al., 2013). Littleis currently understood about different functional roles of lamin-A and lamin-C, although differences in both physical properties,

    such as mobility (Broers et al., 2005), and lamina assembly (Kolbet al., 2011; Pugh et al., 1997) have been reported.

    Stress-responsive regulation of lamin – ‘use it or lose it’To better understand how lamin proteins are actively regulated inresponse to stress, mesenchymal stem cells (MSCs) were cultured

    on type-I-collagen-coated polyacrylamide hydrogels withstiffnesses that were set to mimic the ECM of either brain(0.3 kPa) or pre-calcified bone (40 kPa) (Buxboim et al., 2010b;Swift et al., 2013b). Images of the cultured MSCs showed

    that the nuclear envelopes of cells on soft substrate arewrinkled and relaxed, whereas, on stiff substrate, the nuclei areflattened by stress fibers and appear taut and smooth (Fig. 3B,

    left panel). Accompanying proteomic analyses revealed that thenative fold of A-type lamin is maintained on stiff substrate(specifically, that of the globular Ig-like domain), but that the

    total quantity of A-type lamin protein is upregulated.Furthermore, the extent of phosphorylation at four sites isdecreased on stiff substrate. Phosphorylation is recognized as akey mechanism for modulating the solubility, conformation and

    organization of intermediate filament proteins (Omary et al.,2006), and indeed, lamins are highly phosphorylated during

    Lamin-A increases nuclear viscosity

    L

    ΔP

    Lamin-B(elastic)

    Lamin-A(viscous)

    ViscosityElasticity

    Nuc

    lear

    com

    plia

    nce

    Time (seconds)

    Low LMNA

    MediumLMNA

    High LMNA

    Soft

    Stiff

    Nucleus (GFP-lamin-A)

    Micropipette

    Tim

    e

    Deformationresponse time

    A CB

    Fig. 2. The mechanical role of lamin in the nucleus. (A) Deformations applied by micropipette aspiration were used to quantify nuclear compliance(effectively a measure of ‘softness’, the inverse of stiffness) in a range of nuclei with experimentally altered lamina compositions (achieved, for example, byoverexpressing a GFP–lamin-A fusion construct). Compliance can be calculated over a range of deformation timescales as a function of the micropipettediameter, the extent of deformation (L) and the applied pressure (DP). (B) When a constant deforming pressure was delivered by micropipette over timescales onthe order of seconds, nuclei with low expression of lamin-A (LMNA) were found to be more compliant than those with high expression of lamin-A. (C) Themechanical properties of the lamina can be considered as a combination of elastic (spring-like) and viscous (flowing) properties, which together define the‘deformation response time’, the timescale over which the nuclear shape deforms under force. Nuclei with greater quantities of A-type lamins relative to B-typelamins were found to deform more slowly under stress (Swift et al., 2013b).

    COMMENTARY Journal of Cell Science (2014) 127, 3005–3015 doi:10.1242/jcs.149203

    3008

  • Jour

    nal o

    f Cel

    l Sci

    ence

    normal mitosis, driving the disassembly of the lamina inpreparation for chromosomal separation (Gerace and Blobel,

    1980; Heald and McKeon, 1990). Thus, the response we observeto substrate-induced stress is the converse of this process, withdecreased phosphorylation acting to decrease A-type lamin

    solubility and so strengthening the lamina. On soft substrate, A-type lamins are more extensively phosphorylated, more mobileand so, we hypothesize, more susceptible to turnover (as observedfollowing Akt-mediated phosphorylation of S404 in prelamin-A

    (Bertacchini et al., 2013). These observations hence point to a‘use it or lose it’ dynamic, whereby non-essential A-type lamin is

    eventually degraded. The level of A-type lamins has beenreported to drive the translocation of the lamin-promoting

    transcription factor retinoic acid receptor c (RARG) to thenucleus, pointing to a feedback mechanism by which laminprotein promotes its own transcription (see gene circuit in Swift

    et al., 2013b).There are a number of cases in the literature where cell mechanics

    are regulated by phosphorylation or, by contrast, where externalstresses drive changes in phosphorylation state. As an example of

    the former, changes in the phosphorylation of microtubuleassociated proteins (MAPs) affect the assembly of microtubules

    Cytoplasm

    Alternativesplicing

    miRNA Translation

    Solubility-regulatedby phosphorylation Degradation

    Nuclear laminaAssembly (under stress)

    Disassembly (relaxed)

    LMNA

    Stress-inducedprotein unfolding

    Stress-inducedmembrane crackingTranscription factor andsmall molecule import

    LMNA promoter

    Nucleoplasmicphospho-lamin

    LMNCLamin-ALamin-CTranscription

    Chromatin interactionwith lamina

    Tension

    A-typelamin

    B-typelamin

    Kinase X

    LMNA

    Soft substrate relaxed nuclei

    Stiff substrate tense nuclei

    Ig domain

    TailCoiled coil

    Tension

    Coiled-coil filaments Proteases

    Retinoic-acid-binding factors

    Nucleus

    Wrinkled lamina Smooth lamina

    Turnover ofunstressedfilaments

    A Regulation of A-type lamin protein feeds back on transcription

    B Tension-dependent regulation of protein levels

    Time

    Probe tips used to apply strain

    Fig. 3. Protein regulation as a function of stress. (A) Schematic showing the factors that can regulate the levels of A-type lamin in the cell. LMNAtranscription is promoted by retinoic-acid-binding factors (Okumura et al., 2004a; Olins et al., 2001). The transcript (gray) is alternatively spliced to give rise tolamin-A and lamin-C forms. In some tissues, such as brain, the lamin-A form is suppressed through miRNA (Jung et al., 2012). Mature lamin-A (following post-translational processing) and lamin-C (both shown in red) assemble into the nuclear lamina, although some protein remains mobile in the nucleoplasm (Shimiet al., 2008). Phosphorylation leads to increased solubility, and might precede enzymatic protein turnover. Further stress-dependent pathways have beenreported; stress on the nucleus causes unfolding of the Ig-domain of A-type lamin, and phosphorylation is suppressed under tension (Swift et al., 2013b).Laminopathic nuclei have been shown to have transient membrane defects that allow ingress of transcription factors (De Vos et al., 2011). (B) Left: the nuclei ofMSCs cultured on soft substrate are wrinkled, whereas those in cells on stiff substrate have a smooth stretched appearance, suggestive of greater tension.We have shown that A-type lamin is less phosphorylated under tension – we hypothesize that this might be because interaction with kinases is altered by stress-induced changes in the organization of lamin protein filaments (Swift et al., 2013b). A lower level of phosphorylation reduces the solubility and mobility of laminand thus drives a ‘stress-strengthening’ of the lamina (Swift et al., 2013b). Right: collagen fibrils under tension have been shown to be less susceptible toenzymatic digestion; when a film of fibrillar collagen is subjected to localized stretching, subsequent treatment with a matrix metalloproteinase degrades thecollagen but the fibrils that are under tension remain intact (Flynn et al., 2010). This experiment is another example of multimeric coiled-coil protein assemblybeing enzymatically regulated when subjected to stress, suggesting some generality for the mechanism. Arrowheads indicate the direction of tension.

    COMMENTARY Journal of Cell Science (2014) 127, 3005–3015 doi:10.1242/jcs.149203

    3009

  • Jour

    nal o

    f Cel

    l Sci

    ence

    that have important structural roles in the cytoskeleton (Mateniaand Mandelkow, 2009). Instances of the second case include both

    the extension-dependent phosphorylation of the Cas substratedomain protein, p130Cas (also known as BCAR1), whichtransduces mechanical stretching into biochemical signaling(Sawada et al., 2006), and the phosphorylation of non-muscle

    myosin-IIa (encoded by MYH9), which is inversely related tosubstrate stiffness and acts to regulate cell contractility (Raab et al.,2012). Recent work has shown that the nuclear envelope protein

    emerin is phosphorylated in response to cell contractility, whichconsequently influences nuclear rigidity and, potentially, the wayin which emerin interacts with other nuclear proteins such

    as lamin-A and lamin-C (Guilluy et al., 2014). Furthermore,phosphorylation events could be key to driving the transport ofproteins such as transcription factors to the nucleus, where they

    have crucial regulatory roles (Jans and Hübner, 1996).The concept of ‘stress strengthening’ is also found in other

    biological systems; stretched type I collagen fibrils were found tobe less susceptible to enzymatic degradation (Fig. 3B, right

    panel; Flynn et al., 2010). Although the mode of enzymatic actiondiffers between A-type lamin and type I collagen, both areexamples of polymeric structural proteins that are protected from

    degradation pathways when they are under tension. Theseparallels should also serve as reminders that, although thisreview is primarily focused on stress response within the cell, the

    ECM in living tissue is also being continuously remodeled inresponse to extracellular cues (Xu et al., 2009).

    Mechanotransduction to the nucleus – downstream of ECMand laminLamin is a key component in a system of protein linkages thatallows the transmission of signals from the surroundings of a cell

    into the transcriptional machinery of its nucleus (Fig. 1A;discussed in recent reviews, e.g. Gundersen and Worman, 2013;Rothballer and Kutay, 2013; Simon and Wilson, 2011; Sosa et al.,

    2013). Cell–cell interactions link to the cytoskeleton throughtight junctions and adherens junctions that tether to actin, andthrough desmosome complexes that interact with cytoplasmic

    intermediate filaments, such as keratin (Jamora and Fuchs, 2002).Cell–ECM interactions are mediated by integrins and focaladhesion complexes that bind to cytoplasmic actin (Puklin-Faucher and Sheetz, 2009; Watt and Huck, 2013). The

    appropriately named LINC complex (linker of nucleoskeletonand cytoskeleton) acts as an intermediary between cytoplasmicand nuclear structural proteins; F-actin binds to the nuclear

    envelope components nesprin-1 and nesprin-2, and intermediatefilaments bind to the desmosome protein plectin, which, in turn,binds to nesprin-3. Nesprins can also interact with kinesin and

    dynein complexes to tether to the microtubule network; nesprinsbind to the SUN-domain-containing family of inner nuclearmembrane proteins and these, in turn, bind to the lamina on the

    inside of the nuclear envelope.Lamins engage in extensive protein–protein interactions within

    the nucleus (Wilson and Berk, 2010; Wilson and Foisner, 2010),as emphasized by Wilson and Berk in their review: ‘‘almost all

    characterized [inner nuclear membrane] proteins bind to A- or B-type lamins (or both) directly’’. These interactions includebinding to structural proteins, like actin (Simon et al., 2010),

    and binding to a range of proteins that interact with the nuclearmembrane, including emerin, barrier-to-autointegration factor(BANF; Montes de Oca et al., 2009), lamina-associated

    polypeptide 2 (LAP2, also known as ERBB2IP) and lamin-B

    receptor (LBR; Solovei et al., 2013). Of these, emerin hasattracted considerable recent interest for its role in mediating

    changes in the stiffness of isolated nuclei in response to tensionapplied to nesprin-1 (Guilluy et al., 2014), and its role inmechanosensing through its effect on the translocation of thetranscription factor MKL1 (Ho et al., 2013). Furthermore, some

    transcription factors, such as Oct-1, interact with the laminadirectly (Malhas et al., 2009). Many lamin-binding proteins alsointeract with chromatin, particularly in its silenced heterochromatic

    form (Wagner and Krohne, 2007), and indeed, the lamins havebeen shown to bind to DNA directly (Ludérus et al., 1992;Shoeman and Traub, 1990; Stierlé et al., 2003). This chain of

    interactions thus completes a continuous physical linkage throughwhich deformations can be transmitted from the cell exterior tochromatin (Maniotis et al., 1997). What is missing from this

    picture, however, is an understanding of the mechanism by whichblunt inputs – forces and perturbations acting without microscopiccoherence – can be converted from mechanical to biochemicalsignals to activate individual genes at precise spatial locations

    within the nucleus. Perhaps specificity can be delivered throughchanges in binding, local concentration, conformation andmodification of cofactors or transcription factors.

    As described earlier in this Commentary, mechanical signalsfrom outside the cell are reflected in alterations in proteinconformations, modifications and expression levels, and these can

    broadly affect cell morphology. Taking these changes further, it isof particular interest to understand how external factors drivestem cells to commit to specific lineages, with far-reaching

    implications for therapeutics and tissue engineering. Populationsof MSCs can be expanded in culture in an undifferentiated state,but they can differentiate into mesenchymal lineages, includingfat, cartilage, muscle and bone, in a manner that is dependent on

    external cues, such as the presence of nutrients, growth factorsand cytokines, cell density, spatial constraints and mechanicalforces (Pittenger et al., 1999). It has been shown that cell shape

    influences cell fate by modulating the activity of the smallGTPase RhoA, with round cells being predisposed towardsadipogenesis and well-spread cells more likely to enter the

    osteogenic lineage (McBeath et al., 2004). Here, RhoA was foundto drive commitment to a specific lineage in conjunction with itseffector Rho-associated protein kinase (ROCK), through its effecton cytoskeletal tension.

    Cell fate can be influenced by the stiffness of the ECM (Engleret al., 2006), and we have recently shown that this effect can bemodulated by the nuclear lamina (Fig. 4A; Swift et al., 2013b).

    MSCs cultured on soft hydrogel substrates are biased towardsadipogenesis, but the extent of adipogenesis, as determined bystaining with Oil Red O, is doubled when combined with the

    knockdown of A-type lamins. Likewise, a cooperative effect betweenstiff substrate and lamin-A overexpression enhances osteogenesis.Stiff substrate has been found to drive the translocation of the

    transcription factors RARG and Yes-associated protein 1 (YAP1) tothe nucleus, thus directly regulating the transcription of LMNA andtriggering a differentiation-inducing signaling cascade (Dupont et al.,2011). Culturing MSCs on stiff substrate also increases the protein

    levels of cytoskeletal components, consistent with greater activity ofserum response factor (SRF) (Connelly et al., 2010; Ho et al., 2013),and, in combination with a stiffer nucleus, this further increases

    cytoskeletal tension. Translocation of molecular signal carriers is arecurring motif in mechanosensitive pathways that has been shown tooccur during the transfer of ions and changes in osmotic pressure

    (Finan et al., 2009; Irianto et al., 2013; Kalinowski et al., 2013), as

    COMMENTARY Journal of Cell Science (2014) 127, 3005–3015 doi:10.1242/jcs.149203

    3010

  • Jour

    nal o

    f Cel

    l Sci

    ence

    well as during the migration of protein factors. Protein translocationcould be driven by a change in the concentration of binding sites,(e.g. on A-type lamins or emerin; Ho et al., 2013; Swift et al., 2013b)

    or, conceivably, by protein modifications (e.g. YAP1 nuclearlocalization can be stimulated by phosphorylation of the protein;Murphy et al., 2014). Additionally, transient breakdown of thenuclear envelope in laminopathic cells, perhaps as a consequence of a

    reduced robustness under conditions of mechanical stress, has beenshown to allow the ingress of transcription factors such as RelA (DeVos et al., 2011).

    The ability of lamin and its binding partners to tether to DNAhas led to interest in its role in chromatin organization andregulation (Guelen et al., 2008; Kim et al., 2011; Kind et al.,

    2013; Lund et al., 2013; Meuleman et al., 2013; Zullo et al.,2012). Lamina-associated domains (LADs) located at the nuclearperiphery are typically associated with low levels of gene

    expression, whereas actively transcribed euchromatin is usuallyfound in the nuclear interior. These conserved spatialrelationships within the nucleus give rise to defined‘chromosome territories’ (Cremer and Cremer, 2001; Iyer et al.,

    2012) and to transcriptional hotspots within specific locations(Fraser and Bickmore, 2007). Although chromatin and DNA aregenerally considered to make negligible contributions to overall

    nuclear mechanics (Guilluy et al., 2014; Pajerowski et al., 2007),particular cases are emerging – for example in ESCs passingthrough a metastable transitional state before differentiation –

    where the condensation state of chromatin can becomemechanically significant (Pagliara et al., 2014). It is not yetfully understood which proteins could give rise to mechanically

    responsive locally defined structures and organization within thenucleus, nor how a protein as ubiquitously expressed as lamincould play a part in such specificity. Knowledge in this field willcontinue to improve as new experimental methods and models

    emerge to allow the study of protein-mediated changes inchromatin organization in response to perturbation of the

    system (Shivashankar, 2011; Talwar et al., 2013). However,based on current work, we have hypothesized that the effect oflamin on nuclear mechanics could determine the sensitivity and

    timescale of nuclear reorganization in response to stress (Fig. 4B;Swift et al., 2013b).

    Cell migration is slowed by the nuclear stiffness needed toprotect chromatinAs the nucleus is generally the largest and stiffest organelle, it canbe a limiting factor in the migration of a cell through the three-

    dimensional ECM. This means that the mechanical properties ofthe nucleus can have regulatory roles in certain processes, such asdevelopment, wound healing, hematopoiesis, cancer metastasis

    and others (Fig. 5A, left panel). Studies of migration throughnarrow pores that mimic those in tumor tissue, migration throughwhich requires the deformation of the nucleus, demonstrated a

    dependence on lamina composition – migration was limited whenlamin-A was overexpressed, but was promoted by a ,50%knockdown of the protein (Harada et al., 2014). However, a moreeffective knockdown to ,10% of normal protein expressionlevels was found to cause apoptosis, underscoring the importanceof lamin in providing physical protection to the nucleus (Fig. 5A,middle panel). Consistent with earlier observations that A-type

    lamin and B-type lamin contribute primarily viscous and elasticmechanical properties to nuclei, respectively (Fig. 2), nuclei inwhich high A-type lamin levels dominated the mechanical

    characteristics were observed to recover slowly followingdeformation, maintaining an elongated morphology afteremerging from the pores (Fig. 5A, upper-right panel). By

    contrast, nuclei with dominant levels of elastic B-type laminsrapidly returned to their more spheroid pre-migratory shapesfollowing deformation (Fig. 5A, lower-right panel).

    Cell migration is an important part of the development process,

    and it is possible that the elasticity imparted by lamin-B is neededto allow nuclei to recover from the deformation (typically

    Cell fate is influenced downstream of lamin-A

    LMNA LMNA

    A-type lamin level determinesthe rate of deformation when

    the nucleus is stressed

    New chromatinand lamina contacts

    A B

    Soft matrix Stiff matrixLow A-type lamin High A-type lamin

    Balled-up cells Spread cells, stress fibersCytoplasmic RARG and YAP1 Nucleoplasmic RARG and YAP1

    Adipogenesis Osteogenesis

    Chromosome territories

    Interchromatin contacts

    Lamina-chromatin contacts

    Tension

    Fig. 4. Decisions of cell fate downstream of the regulation of A-type lamin. (A) MSCs cultured on soft and stiff substrates take on different phenotypesand are biased towards alternative cell fates (Engler et al., 2006). On soft substrate, MSCs are rounded and exhibit small nuclear and cellular areas, and thenuclear lamina is thinned (thin red nuclear outline) by a stress-sensitive phosphorylation-feedback mechanism (Fig. 3B, left panel; Swift et al., 2013b). Thetranscription factors RARG and YAP1 (Dupont et al., 2011) remain in the cytoplasm, and adipogenic cell fate is preferred. On stiff substrate, cells spreadextensively, and the nuclei are pinned down by well-developed stress fibers (gray). A-type lamin is less phosphorylated under strain, thus strengthening thelamina (thick red nuclear outline). RARG also translocates to the nucleus, increasing LMNA transcription. Activity of the transcription factor SRF(downstream of A-type lamin) increases the expression of cytoskeletal components (Ho et al., 2013). Under these conditions, YAP1 translocates to the nucleusand cells are more likely to undergo osteogenesis. On both soft and stiff substrates, the effects of ECM elasticity and the levels of lamin cooperate to enhancedifferentiation; knockdown of A-type lamin on soft substrate leads to more adipogenesis, and lamin-A overexpression on stiff substrate leads to moreosteogenesis. (B) Transcriptional activity is believed to be regulated by conserved interchromatin contacts that give rise to the spatial ordering of chromosomes(chromosome territories shown here in different colors, Cremer and Cremer, 2001). A-type lamin can interact with DNA directly (lamina–chromatin contacts)or through protein intermediaries (Simon and Wilson, 2011), but could have additional regulatory roles by mechanically determining the extent and rate that thenucleus deforms under tension, a process that could lead to the formation of altered interchromatin and lamina–chromatin contacts. Arrowheads indicate thedirection of tension.

    COMMENTARY Journal of Cell Science (2014) 127, 3005–3015 doi:10.1242/jcs.149203

    3011

  • Jour

    nal o

    f Cel

    l Sci

    ence

    elongation) that occurs during migration, perhaps explaining whythe brain fails to develop in lamin-B-knockout mice (Coffinieret al., 2011; Jung et al., 2013; Kim et al., 2011). Neutrophilic cells

    also have very low levels of nucleoskeletal proteins to allow theirdeformation as they squeeze into confined spaces (Olins et al.,2009; Rowat et al., 2013), and indeed, the composition of thenuclear lamina is continuously regulated during hematopoiesis

    (Fig. 5B; Shin et al., 2013). We hypothesize that bydownregulating components of the lamina, white blood cellscompromise their robustness in favor of mobility, and that this

    contributes to the short lifetimes of many of these cells incirculation. Cancer metastasis is an equally complex process thatdepends on factors including the remodeling of the ECM and the

    capacity of the nucleus to deform (Harada et al., 2014; Wolfet al., 2013). Other work has shown that the ability of myosin-IIto deform the nucleus can be a decisive factor in limiting glioma

    migration into brain tissue (Beadle et al., 2008; Ivkovic et al.,2012), but cancer cells in general show no universal lamina

    phenotype (reviewed in Foster et al., 2010). Although low levelsof A-type lamin have been correlated with increased reoccurrenceof colon cancers (Belt et al., 2011), A-type lamin was also found

    to be upregulated in certain skin and ovarian cancers (Hudsonet al., 2007; Tilli et al., 2003), and higher expression of theselamin proteins has been associated with better clinical outcomesin breast cancer (Wazir et al., 2013). Our own studies of tumor

    expansion in mouse flank have been suggestive of an associationbetween moderately reduced levels of A-type lamin and increasedinvasiveness into the surrounding tissue, but a more complex

    relationship between the level of A-type lamin and clinicalprognosis might be explained by the tenuous balance between theeffect of the lamina on nuclear deformability compared with that

    on cell survival.

    Conclusions and prospectsWe have sought to emphasize the underlying importance ofnuclear mechanics in the context of tissue function, considering

    A Nuclear mechanics influence 3D cell migration

    Seru

    m

    ECM

    Cell mass

    Nuclei

    Bottom

    Top

    10 μμm

    10 μμm

    Low lamin A:B ratio

    High lamin A:B ratio

    Hematopoieticstem cellprogenitor

    Whiteblood cell

    Redblood cell Platelets

    Enucleation

    MegakaryocyteErythroidprogenitor

    Granulocyte,monocyte and

    lymphocye

    Lamin-A expression

    Knockdown Overexpression

    Mic

    ropo

    re m

    igra

    tion

    Very low lamin-A:apoptosis

    Lamin-A KD:increased migration

    Lamin-A OE:impeded migration

    WT

    Bone marrow:higher lamins

    Peripheral blood:low lamins

    Shear

    B Lamina composition regulates hematopoiesis

    Fig. 5. The influence of the mechanical properties of the nucleus on cell migration. (A) Left: as the largest and stiffest organelle in the cell, the nucleus canact as an ‘anchor’, preventing cell movement through the ECM or into the surrounding vasculature. Middle: to model migration through the ECM, cells areinduced to pass through 3-mm pores, a diameter sufficiently small to require deformation of the nucleus (inset). Lamin-A overexpression (OE) inhibits migration,whereas knockdown (KD) increases migration to five times that of wild-type (WT) cells; however, highly effective knockdown leads to substantial apoptosis.Thus, extremely low or high levels of A-type lamin are unfavorable for cell migration, an observation that potentially impacts upon the understanding of processessuch as cell migration during development and cancer metastasis. Right: lamin-A-rich nuclei (upper panel) show a persistently elongated morphology uponemerging from the pores (yellow arrowheads), whereas lamin-B-rich nuclei (lower panel) rapidly recover their shape (Fig. 5A adapted from Harada et al., 2014.Originally published in The Journal of Cell Biology, doi: 10.1083/jcb.201308029). (B) The effect of lamina composition on nuclear deformability duringhematopoiesis. Stem cells that are retained in the marrow niche have higher lamin levels than differentiated blood lineages (Shin et al., 2013). A downregulationof nuclear cytoskeletal components in granulocytes, for example, ostensibly makes the cells better suited for passage through narrow blood vessels, but the lackof nuclear stability might contribute to their relatively short circulation times (Olins et al., 2009).

    COMMENTARY Journal of Cell Science (2014) 127, 3005–3015 doi:10.1242/jcs.149203

    3012

  • Jour

    nal o

    f Cel

    l Sci

    ence

    how they determine the protective properties of the lamina,influence cell fate and also regulate cell migration. In

    understanding that one of the key functions of lamins is toensure that the mechanical properties of the cell meet thedemands of a tissue – either directly or indirectly, by drivingbroader changes with regard to cell fate – we are essentially

    allocating lamin a role in stress response (Zuela et al., 2012). Theresponse to cellular stress is classically thought of in terms of howcells mitigate heat shock, which can otherwise result in high

    levels of unfolded proteins (Hartl et al., 2011), but mechanicalstress can also cause protein unfolding with associated loss offunction (Swift et al., 2013b). Another recent review highlighted

    the similarities between cellular responses that involve theexpression of heat-shock proteins and those that drive structuralintermediate filament proteins (Toivola et al., 2010). Just as the

    lamina responds to mechanical inputs from the ECM, mechanicalstretching is sufficient to induce the expression of the molecularchaperone heat-shock protein HSP70 (Silver and Noble, 2012).However, we are still a long way from fully understanding

    cellular protection mechanisms and the way in which stressresponse pathways affect the regulation of structural featureswithin the cell. Nonetheless, it is apparent that nuclear biophysics

    has an important role to play.

    AcknowledgementsWe thank our colleagues P. C. Dave P. Dingal, Irena L. Ivanovska and StephanieMajkut at the University of Pennsylvania for helpful comments.

    Competing interestsThe authors declare no competing interests.

    FundingWe appreciate support from the U.S. National Institutes of Health, the U.S.National Science Foundation and the University of Pennsylvania’s researchcenters (Materials Research Science and Engineering; Nano Science andEngineering; Nano/Bio Interface). Deposited in PMC for release after 12 months.

    ReferencesAletras, A. H., Ding, S., Balaban, R. S. andWen, H. (1999). DENSE: displacementencoding with stimulated echoes in cardiac functional MRI. J. Magn. Reson. 137,247-252.

    Beadle, C., Assanah, M. C., Monzo, P., Vallee, R., Rosenfeld, S. S. and Canoll,P. (2008). The role of myosin II in glioma invasion of the brain. Mol. Biol. Cell 19,3357-3368.

    Belt, E. J. T., Fijneman, R. J. A., van den Berg, E. G., Bril, H., Delis-van Diemen,P. M., Tijssen, M., van Essen, H. F., de Lange-de Klerk, E. S., Beliën, J. A.,Stockmann, H. B. et al. (2011). Loss of lamin A/C expression in stage II and IIIcolon cancer is associated with disease recurrence.Eur. J. Cancer 47, 1837-1845.

    Bertacchini, J., Beretti, F., Cenni, V., Guida, M., Gibellini, F., Mediani, L., Marin,O., Maraldi, N. M., de Pol, A., Lattanzi, G. et al. (2013). The protein kinase Akt/PKB regulates both prelamin A degradation and Lmna gene expression. FASEBJ. 27, 2145-2155.

    Bonne, G., Di Barletta, M. R., Varnous, S., Bécane, H. M., Hammouda, E. H.,Merlini, L., Muntoni, F., Greenberg, C. R., Gary, F., Urtizberea, J. A. et al.(1999). Mutations in the gene encoding lamin A/C cause autosomal dominantEmery-Dreifuss muscular dystrophy. Nat. Genet. 21, 285-288.

    Broers, J. L. V., Machiels, B. M., Kuijpers, H. J. H., Smedts, F., van denKieboom, R., Raymond, Y. and Ramaekers, F. C. (1997). A- and B-type laminsare differentially expressed in normal human tissues. Histochem. Cell Biol. 107,505-517.

    Broers, J. L. V., Peeters, E. A. G., Kuijpers, H. J. H., Endert, J., Bouten, C. V.C., Oomens, C. W. J., Baaijens, F. P. T. and Ramaekers, F. C. S. (2004).Decreased mechanical stiffness in LMNA-/- cells is caused by defective nucleo-cytoskeletal integrity: implications for the development of laminopathies. Hum.Mol. Genet. 13, 2567-2580.

    Broers, J. L. V., Kuijpers, H. J. H., Ostlund, C., Worman, H. J., Endert, J. andRamaekers, F. C. S. (2005). Both lamin A and lamin C mutations cause laminainstability as well as loss of internal nuclear lamin organization. Exp. Cell Res.304, 582-592.

    Butin-Israeli, V., Adam, S. A., Goldman, A. E. and Goldman, R. D. (2012).Nuclear lamin functions and disease. Trends Genet. 28, 464-471.

    Buxboim, A., Ivanovska, I. L. and Discher, D. E. (2010a). Matrix elasticity,cytoskeletal forces and physics of the nucleus: how deeply do cells ‘feel’ outsideand in? J. Cell Sci. 123, 297-308.

    Buxboim, A., Rajagopal, K., Brown, A. E. X. and Discher, D. E. (2010b). Howdeeply cells feel: methods for thin gels. J. Phys. Condens. Matter 22, 194116.

    Cance, W. G., Chaudhary, N., Worman, H. J., Blobel, G. and Cordoncardo, C.(1992). Expression of the nuclear lamins in normal and neoplastic humantissues. J. Exp. Clin. Cancer Res. 11, 233-246.

    Chan, D. D. and Neu, C. P. (2012). Transient and microscale deformations andstrains measured under exogenous loading by noninvasive magneticresonance. PLoS ONE 7, e33463.

    Coffinier, C., Jung, H. J., Nobumori, C., Chang, S., Tu, Y., Barnes, R. H., 2nd,Yoshinaga, Y., de Jong, P. J., Vergnes, L., Reue, K. et al. (2011). Deficienciesin lamin B1 and lamin B2 cause neurodevelopmental defects and distinctnuclear shape abnormalities in neurons. Mol. Biol. Cell 22, 4683-4693.

    Connelly, J. T., Gautrot, J. E., Trappmann, B., Tan, D. W. M., Donati, G., Huck,W. T. S. and Watt, F. M. (2010). Actin and serum response factor transducephysical cues from the microenvironment to regulate epidermal stem cell fatedecisions. Nat. Cell Biol. 12, 711-718.

    Cortese, R., Eckhardt, F., Volleth, M., Wehnert, M., Koelsch, U., Wieacker, P.and Brune, T. (2007). The retinol acid receptor B gene is hypermethylated inpatients with familial partial lipodystrophy. J. Mol. Endocrinol. 38, 663-671.

    Cremer, T. and Cremer, C. (2001). Chromosome territories, nuclear architectureand gene regulation in mammalian cells. Nat. Rev. Genet. 2, 292-301.

    Dahl, K. N., Engler, A. J., Pajerowski, J. D. and Discher, D. E. (2005). Power-lawrheology of isolated nuclei with deformation mapping of nuclear substructures.Biophys. J. 89, 2855-2864.

    Davidson, P. M. and Lammerding, J. (2014). Broken nuclei – lamins, nuclearmechanics, and disease. Trends Cell Biol. 24, 247-256.

    De Sandre-Giovannoli, A., Chaouch, M., Kozlov, S., Vallat, J. M., Tazir, M.,Kassouri, N., Szepetowski, P., Hammadouche, T., Vandenberghe, A., Stewart,C. L. et al. (2002). Homozygous defects in LMNA, encoding lamin A/C nuclear-envelope proteins, cause autosomal recessive axonal neuropathy in human(Charcot-Marie-Tooth disorder type 2) andmouse.Am. J. Hum. Genet. 70, 726-736.

    De Vos, W. H., Houben, F., Kamps, M., Malhas, A., Verheyen, F., Cox, J.,Manders, E. M. M., Verstraeten, V. L., van Steensel, M. A. M., Marcelis, C. L.M. et al. (2011). Repetitive disruptions of the nuclear envelope invoke temporaryloss of cellular compartmentalization in laminopathies. Hum. Mol. Genet. 20,4175-4186.

    Dechat, T., Adam, S. A., Taimen, P., Shimi, T. and Goldman, R. D. (2010).Nuclear lamins. Cold Spring Harb. Perspect. Biol. 2, a000547.

    Discher, D. E., Janmey, P. and Wang, Y. L. (2005). Tissue cells feel and respondto the stiffness of their substrate. Science 310, 1139-1143.

    Dittmer, T. A. and Misteli, T. (2011). The lamin protein family. Genome Biol. 12,222.

    Dupont, S., Morsut, L., Aragona, M., Enzo, E., Giulitti, S., Cordenonsi, M.,Zanconato, F., Le Digabel, J., Forcato, M., Bicciato, S. et al. (2011). Role ofYAP/TAZ in mechanotransduction. Nature 474, 179-183.

    Eckersley-Maslin, M. A., Bergmann, J. H., Lazar, Z. and Spector, D. L. (2013).Lamin A/C is expressed in pluripotent mouse embryonic stem cells. Nucleus 4,53-60.

    Eden, E., Geva-Zatorsky, N., Issaeva, I., Cohen, A., Dekel, E., Danon, T.,Cohen, L., Mayo, A. and Alon, U. (2011). Proteome half-life dynamics in livinghuman cells. Science 331, 764-768.

    Engler, A. J., Griffin, M. A., Sen, S., Bönnemann, C. G., Sweeney, H. L. andDischer, D. E. (2004). Myotubes differentiate optimally on substrates withtissue-like stiffness: pathological implications for soft or stiff microenvironments.J. Cell Biol. 166, 877-887.

    Engler, A. J., Sen, S., Sweeney, H. L. and Discher, D. E. (2006). Matrix elasticitydirects stem cell lineage specification. Cell 126, 677-689.

    Fatkin, D., MacRae, C., Sasaki, T., Wolff, M. R., Porcu, M., Frenneaux, M.,Atherton, J., Vidaillet, H. J., Jr, Spudich, S., De Girolami, U. et al. (1999).Missense mutations in the rod domain of the lamin A/C gene as causes of dilatedcardiomyopathy and conduction-system disease.N. Engl. J. Med. 341, 1715-1724.

    Finan, J. D., Chalut, K. J., Wax, A. and Guilak, F. (2009). Nonlinear osmoticproperties of the cell nucleus. Ann. Biomed. Eng. 37, 477-491.

    Flynn, B. P., Bhole, A. P., Saeidi, N., Liles, M., Dimarzio, C. A. and Ruberti, J. W.(2010). Mechanical strain stabilizes reconstituted collagen fibrils againstenzymatic degradation by mammalian collagenase matrix metalloproteinase 8(MMP-8). PLoS ONE 5, e12337.

    Foster, C. R., Przyborski, S. A., Wilson, R. G. and Hutchison, C. J. (2010).Lamins as cancer biomarkers. Biochem. Soc. Trans. 38, 297-300.

    Fraser, P. and Bickmore, W. (2007). Nuclear organization of the genome and thepotential for gene regulation. Nature 447, 413-417.

    Frost, H. M. (1987). Bone ‘‘mass’’ and the ‘‘mechanostat’’: a proposal. Anat. Rec.219, 1-9.

    Gardel, M. L., Shin, J. H., MacKintosh, F. C., Mahadevan, L., Matsudaira, P.and Weitz, D. A. (2004). Elastic behavior of cross-linked and bundled actinnetworks. Science 304, 1301-1305.

    Gerace, L. and Blobel, G. (1980). The nuclear envelope lamina is reversiblydepolymerized during mitosis. Cell 19, 277-287.

    Guelen, L., Pagie, L., Brasset, E., Meuleman, W., Faza, M. B., Talhout, W.,Eussen, B. H., de Klein, A., Wessels, L., de Laat, W. et al. (2008). Domainorganization of human chromosomes revealed by mapping of nuclear laminainteractions. Nature 453, 948-951.

    Guilak, F., Ratcliffe, A. and Mow, V. C. (1995). Chondrocyte deformation andlocal tissue strain in articular cartilage: a confocal microscopy study. J. Orthop.Res. 13, 410-421.

    COMMENTARY Journal of Cell Science (2014) 127, 3005–3015 doi:10.1242/jcs.149203

    3013

    http://dx.doi.org/10.1006/jmre.1998.1676http://dx.doi.org/10.1006/jmre.1998.1676http://dx.doi.org/10.1006/jmre.1998.1676http://dx.doi.org/10.1091/mbc.E08-03-0319http://dx.doi.org/10.1091/mbc.E08-03-0319http://dx.doi.org/10.1091/mbc.E08-03-0319http://dx.doi.org/10.1016/j.ejca.2011.04.025http://dx.doi.org/10.1016/j.ejca.2011.04.025http://dx.doi.org/10.1016/j.ejca.2011.04.025http://dx.doi.org/10.1016/j.ejca.2011.04.025http://dx.doi.org/10.1096/fj.12-218214http://dx.doi.org/10.1096/fj.12-218214http://dx.doi.org/10.1096/fj.12-218214http://dx.doi.org/10.1096/fj.12-218214http://dx.doi.org/10.1038/6799http://dx.doi.org/10.1038/6799http://dx.doi.org/10.1038/6799http://dx.doi.org/10.1038/6799http://dx.doi.org/10.1007/s004180050138http://dx.doi.org/10.1007/s004180050138http://dx.doi.org/10.1007/s004180050138http://dx.doi.org/10.1007/s004180050138http://dx.doi.org/10.1093/hmg/ddh295http://dx.doi.org/10.1093/hmg/ddh295http://dx.doi.org/10.1093/hmg/ddh295http://dx.doi.org/10.1093/hmg/ddh295http://dx.doi.org/10.1093/hmg/ddh295http://dx.doi.org/10.1016/j.yexcr.2004.11.020http://dx.doi.org/10.1016/j.yexcr.2004.11.020http://dx.doi.org/10.1016/j.yexcr.2004.11.020http://dx.doi.org/10.1016/j.yexcr.2004.11.020http://dx.doi.org/10.1016/j.tig.2012.06.001http://dx.doi.org/10.1016/j.tig.2012.06.001http://dx.doi.org/10.1242/jcs.041186http://dx.doi.org/10.1242/jcs.041186http://dx.doi.org/10.1242/jcs.041186http://dx.doi.org/10.1088/0953-8984/22/19/194116http://dx.doi.org/10.1088/0953-8984/22/19/194116http://dx.doi.org/10.1371/journal.pone.0033463http://dx.doi.org/10.1371/journal.pone.0033463http://dx.doi.org/10.1371/journal.pone.0033463http://dx.doi.org/10.1091/mbc.E11-06-0504http://dx.doi.org/10.1091/mbc.E11-06-0504http://dx.doi.org/10.1091/mbc.E11-06-0504http://dx.doi.org/10.1091/mbc.E11-06-0504http://dx.doi.org/10.1038/ncb2074http://dx.doi.org/10.1038/ncb2074http://dx.doi.org/10.1038/ncb2074http://dx.doi.org/10.1038/ncb2074http://dx.doi.org/10.1677/JME-07-0035http://dx.doi.org/10.1677/JME-07-0035http://dx.doi.org/10.1677/JME-07-0035http://dx.doi.org/10.1038/35066075http://dx.doi.org/10.1038/35066075http://dx.doi.org/10.1529/biophysj.105.062554http://dx.doi.org/10.1529/biophysj.105.062554http://dx.doi.org/10.1529/biophysj.105.062554http://dx.doi.org/10.1086/339274http://dx.doi.org/10.1086/339274http://dx.doi.org/10.1086/339274http://dx.doi.org/10.1086/339274http://dx.doi.org/10.1086/339274http://dx.doi.org/10.1093/hmg/ddr344http://dx.doi.org/10.1093/hmg/ddr344http://dx.doi.org/10.1093/hmg/ddr344http://dx.doi.org/10.1093/hmg/ddr344http://dx.doi.org/10.1093/hmg/ddr344http://dx.doi.org/10.1101/cshperspect.a000547http://dx.doi.org/10.1101/cshperspect.a000547http://dx.doi.org/10.1126/science.1116995http://dx.doi.org/10.1126/science.1116995http://dx.doi.org/10.1186/gb-2011-12-5-222http://dx.doi.org/10.1186/gb-2011-12-5-222http://dx.doi.org/10.1038/nature10137http://dx.doi.org/10.1038/nature10137http://dx.doi.org/10.1038/nature10137http://dx.doi.org/10.4161/nucl.23384http://dx.doi.org/10.4161/nucl.23384http://dx.doi.org/10.4161/nucl.23384http://dx.doi.org/10.1126/science.1199784http://dx.doi.org/10.1126/science.1199784http://dx.doi.org/10.1126/science.1199784http://dx.doi.org/10.1083/jcb.200405004http://dx.doi.org/10.1083/jcb.200405004http://dx.doi.org/10.1083/jcb.200405004http://dx.doi.org/10.1083/jcb.200405004http://dx.doi.org/10.1016/j.cell.2006.06.044http://dx.doi.org/10.1016/j.cell.2006.06.044http://dx.doi.org/10.1056/NEJM199912023412302http://dx.doi.org/10.1056/NEJM199912023412302http://dx.doi.org/10.1056/NEJM199912023412302http://dx.doi.org/10.1056/NEJM199912023412302http://dx.doi.org/10.1007/s10439-008-9618-5http://dx.doi.org/10.1007/s10439-008-9618-5http://dx.doi.org/10.1371/journal.pone.0012337http://dx.doi.org/10.1371/journal.pone.0012337http://dx.doi.org/10.1371/journal.pone.0012337http://dx.doi.org/10.1371/journal.pone.0012337http://dx.doi.org/10.1042/BST0380297http://dx.doi.org/10.1042/BST0380297http://dx.doi.org/10.1038/nature05916http://dx.doi.org/10.1038/nature05916http://dx.doi.org/10.1002/ar.1092190104http://dx.doi.org/10.1002/ar.1092190104http://dx.doi.org/10.1126/science.1095087http://dx.doi.org/10.1126/science.1095087http://dx.doi.org/10.1126/science.1095087http://dx.doi.org/10.1016/0092-8674(80)90409-2http://dx.doi.org/10.1016/0092-8674(80)90409-2http://dx.doi.org/10.1038/nature06947http://dx.doi.org/10.1038/nature06947http://dx.doi.org/10.1038/nature06947http://dx.doi.org/10.1038/nature06947http://dx.doi.org/10.1002/jor.1100130315http://dx.doi.org/10.1002/jor.1100130315http://dx.doi.org/10.1002/jor.1100130315

  • Jour

    nal o

    f Cel

    l Sci

    ence

    Guilluy, C., Osborne, L. D., Van Landeghem, L., Sharek, L., Superfine, R.,Garcia-Mata, R. and Burridge, K. (2014). Isolated nuclei adapt to force andreveal a mechanotransduction pathway in the nucleus. Nat. Cell Biol. 16, 376-381.

    Gundersen, G. G. and Worman, H. J. (2013). Nuclear positioning. Cell 152,1376-1389.

    Hadjipanayi, E., Mudera, V. and Brown, R. A. (2009a). Close dependence offibroblast proliferation on collagen scaffold matrix stiffness. J. Tissue Eng.Regen. Med. 3, 77-84.

    Hadjipanayi, E., Mudera, V. and Brown, R. A. (2009b). Guiding cell migration in3D: a collagen matrix with graded directional stiffness. Cell Motil. Cytoskeleton66, 121-128.

    Harada, T., Swift, J., Irianto, J., Shin, J.-W., Spinler, K. R., Athirasala, A.,Diegmiller, R., Dingal, P. C. D. P., Ivanovska, I. L. and Discher, D. E. (2014).Nuclear lamin stiffness is a barrier to 3D migration, but softness can limitsurvival. J. Cell Biol. 204, 669-682.

    Hartl, F. U., Bracher, A. and Hayer-Hartl, M. (2011). Molecular chaperones inprotein folding and proteostasis. Nature 475, 324-332.

    Heald, R. and McKeon, F. (1990). Mutations of phosphorylation sites in lamin Athat prevent nuclear lamina disassembly in mitosis. Cell 61, 579-589.

    Hegele, R. A., Cao, H., Huff, M. W. and Anderson, C. M. (2000). LMNA R482Qmutation in partial lipodystrophy associated with reduced plasma leptinconcentration. J. Clin. Endocrinol. Metab. 85, 3089-3093.

    Henderson, J. T., Shannon, G., Veress, A. I. and Neu, C. P. (2013). Directmeasurement of intranuclear strain distributions and RNA synthesis in singlecells embedded within native tissue. Biophys. J. 105, 2252-2261.

    Herrmann, H., Strelkov, S. V., Burkhard, P. and Aebi, U. (2009). Intermediatefilaments: primary determinants of cell architecture and plasticity. J. Clin. Invest.119, 1772-1783.

    Ho, C. Y. and Lammerding, J. (2012). Lamins at a glance. J. Cell Sci. 125, 2087-2093.

    Ho, C. Y., Jaalouk, D. E., Vartiainen, M. K. and Lammerding, J. (2013). Lamin A/C and emerin regulate MKL1-SRF activity by modulating actin dynamics. Nature497, 507-511.

    Hudson, M. E., Pozdnyakova, I., Haines, K., Mor, G. and Snyder, M. (2007).Identification of differentially expressed proteins in ovarian cancer using high-density protein microarrays. Proc. Natl. Acad. Sci. USA 104, 17494-17499.

    Irianto, J., Swift, J., Martins, R. P., McPhail, G. D., Knight, M. M., Discher, D. E.and Lee, D. A. (2013). Osmotic challenge drives rapid and reversible chromatincondensation in chondrocytes. Biophys. J. 104, 759-769.

    Isermann, P. and Lammerding, J. (2013). Nuclear mechanics andmechanotransduction in health and disease. Curr. Biol. 23, R1113-R1121.

    Ivkovic, S., Beadle, C., Noticewala, S., Massey, S. C., Swanson, K. R., Toro, L. N.,Bresnick, A. R., Canoll, P. and Rosenfeld, S. S. (2012). Direct inhibition ofmyosin II effectively blocks glioma invasion in the presence of multiple motogens.Mol. Biol. Cell 23, 533-542.

    Iyer, K. V., Maharana, S., Gupta, S., Libchaber, A., Tlusty, T. andShivashankar, G. V. (2012). Modeling and experimental methods to probethe link between global transcription and spatial organization of chromosomes.PLoS ONE 7, e46628.

    Jahn, D., Schramm, S., Schnölzer, M., Heilmann, C. J., de Koster, C. G.,Schütz, W., Benavente, R. and Alsheimer, M. (2012). A truncated lamin A inthe Lmna -/- mouse line: implications for the understanding of laminopathies.Nucleus 3, 463-474.

    Jamora, C. and Fuchs, E. (2002). Intercellular adhesion, signalling and thecytoskeleton. Nat. Cell Biol. 4, E101-E108.

    Jans, D. A. and Hübner, S. (1996). Regulation of protein transport to the nucleus:central role of phosphorylation. Physiol. Rev. 76, 651-685.

    Jung, H. J., Coffinier, C., Choe, Y., Beigneux, A. P., Davies, B. S. J., Yang, S. H.,Barnes, R. H., 2nd, Hong, J., Sun, T., Pleasure, S. J. et al. (2012). Regulation ofprelamin A but not lamin C by miR-9, a brain-specific microRNA. Proc. Natl. Acad.Sci. USA 109, E423-E431.

    Jung, H. J., Lee, J. M., Yang, S. H., Young, S. G. and Fong, L. G. (2013).Nuclear lamins in the brain - new insights into function and regulation. Mol.Neurobiol. 47, 290-301.

    Kalinowski, A., Qin, Z., Coffey, K., Kodali, R., Buehler, M. J., Lösche, M. andDahl, K. N. (2013). Calcium causes a conformational change in lamin A taildomain that promotes farnesyl-mediated membrane association. Biophys. J. 104,2246-2253.

    Kim, Y., Sharov, A. A., McDole, K., Cheng, M., Hao, H., Fan, C. M., Gaiano, N.,Ko, M. S. H. and Zheng, Y. (2011). Mouse B-type lamins are required for properorganogenesis but not by embryonic stem cells. Science 334, 1706-1710.

    Kim, Y., Zheng, X. and Zheng, Y. (2013). Proliferation and differentiation ofmouse embryonic stem cells lacking all lamins. Cell Res. 23, 1420-1423.

    Kind, J., Pagie, L., Ortabozkoyun, H., Boyle, S., de Vries, S. S., Janssen, H.,Amendola, M., Nolen, L. D., Bickmore, W. A. and van Steensel, B. (2013).Single-cell dynamics of genome-nuclear lamina interactions. Cell 153, 178-192.

    Klein, E. A., Yin, L., Kothapalli, D., Castagnino, P., Byfield, F. J., Xu, T.,Levental, I., Hawthorne, E., Janmey, P. A. and Assoian, R. K. (2009). Cell-cycle control by physiological matrix elasticity and in vivo tissue stiffening. Curr.Biol. 19, 1511-1518.

    Kolb, T., Maass, K., Hergt, M., Aebi, U. and Herrmann, H. (2011). Lamin A andlamin C form homodimers and coexist in higher complex forms both in thenucleoplasmic fraction and in the lamina of cultured human cells. Nucleus 2,425-433.

    Krohne, G., Dabauvalle, M. C. and Franke, W. W. (1981). Cell type-specificdifferences in protein composition of nuclear pore complex-lamina structures inoocytes and erythrocytes of Xenopus laevis. J. Mol. Biol. 151, 121-141.

    Kubben, N., Voncken, J. W., Konings, G., van Weeghel, M., van denHoogenhof, M. M. G., Gijbels, M., van Erk, A., Schoonderwoerd, K., vanden Bosch, B., Dahlmans, V. et al. (2011). Post-natal myogenic andadipogenic developmental: defects and metabolic impairment upon loss of A-type lamins. Nucleus 2, 195-207.

    Lammerding, J., Fong, L. G., Ji, J. Y., Reue, K., Stewart, C. L., Young, S. G.and Lee, R. T. (2006). Lamins A and C but not lamin B1 regulate nuclearmechanics. J. Biol. Chem. 281, 25768-25780.

    Lehner, C. F., Stick, R., Eppenberger, H. M. and Nigg, E. A. (1987). Differentialexpression of nuclear lamin proteins during chicken development. J. Cell Biol.105, 577-587.

    Lo, C. M., Wang, H. B., Dembo, M. and Wang, Y. L. (2000). Cell movement isguided by the rigidity of the substrate. Biophys. J. 79, 144-152.

    Ludérus, M. E. E., de Graaf, A., Mattia, E., den Blaauwen, J. L., Grande, M. A.,de Jong, L. and van Driel, R. (1992). Binding of matrix attachment regions tolamin B1. Cell 70, 949-959.

    Lund, E., Oldenburg, A. R., Delbarre, E., Freberg, C. T., Duband-Goulet, I.,Eskeland, R., Buendia, B. and Collas, P. (2013). Lamin A/C-promoterinteractions specify chromatin state-dependent transcription outcomes.Genome Res. 23, 1580-1589.

    Majkut, S., Idema, T., Swift, J., Krieger, C., Liu, A. and Discher, D. E. (2013).Heart-specific stiffening in early embryos parallels matrix and myosin expressionto optimize beating. Curr. Biol. 23, 2434-2439.

    Malhas, A. N., Lee, C. F. and Vaux, D. J. (2009). Lamin B1 controls oxidativestress responses via Oct-1. J. Cell Biol. 184, 45-55.

    Maniotis, A. J., Chen, C. S. and Ingber, D. E. (1997). Demonstration of mechanicalconnections between integrins, cytoskeletal filaments, and nucleoplasm thatstabilize nuclear structure. Proc. Natl. Acad. Sci. USA 94, 849-854.

    Matenia, D. and Mandelkow, E. M. (2009). The tau of MARK: a polarized view ofthe cytoskeleton. Trends Biochem. Sci. 34, 332-342.

    McBeath, R., Pirone, D. M., Nelson, C. M., Bhadriraju, K. and Chen, C. S.(2004). Cell shape, cytoskeletal tension, and RhoA regulate stem cell lineagecommitment. Dev. Cell 6, 483-495.

    Merideth, M. A., Gordon, L. B., Clauss, S., Sachdev, V., Smith, A. C. M., Perry,M. B., Brewer, C. C., Zalewski, C., Kim, H. J., Solomon, B. et al. (2008). Phenotypeand course of Hutchinson-Gilford progeria syndrome.N. Engl. J. Med. 358, 592-604.

    Meuleman, W., Peric-Hupkes, D., Kind, J., Beaudry, J. B., Pagie, L., Kellis, M.,Reinders, M., Wessels, L. and van Steensel, B. (2013). Constitutive nuclearlamina-genome interactions are highly conserved and associated with A/T-richsequence. Genome Res. 23, 270-280.

    Montes de Oca, R., Shoemaker, C. J., Gucek, M., Cole, R. N. and Wilson, K. L.(2009). Barrier-to-autointegration factor proteome reveals chromatin-regulatorypartners. PLoS ONE 4, e7050.

    Murphy, A. J., Pierce, J., de Caestecker, C., Libes, J., Neblett, D., deCaestecker, M., Perantoni, A. O., Tanigawa, S., Anderson, J. R., Dome, J. S.et al. (2014). Aberrant activation, nuclear localization, and phosphorylation ofYes-associated protein-1 in the embryonic kidney and Wilms tumor. Pediatr.Blood Cancer 61, 198-205.

    Okumura, K., Hosoe, Y. and Nakajima, N. (2004a). c-Jun and Sp1 family arecritical for retinoic acid induction of the lamin A/C retinoic acid-responsiveelement. Biochem. Biophys. Res. Commun. 320, 487-492.

    Okumura, K., Hosoe, Y. and Nakajima, N. (2004b). Zic1 is a transcriptionalrepressor through the lamin A/C promoter and has an intrinsic repressivedomain. J. Health Sci. 50, 423-427.

    Olins, A. L., Herrmann, H., Lichter, P., Kratzmeier, M., Doenecke, D. and Olins,D. E. (2001). Nuclear envelope and chromatin compositional differencescomparing undifferentiated and retinoic acid- and phorbol ester-treated HL-60cells. Exp. Cell Res. 268, 115-127.

    Olins, A. L., Hoang, T. V., Zwerger, M., Herrmann, H., Zentgraf, H., Noegel, A. A.,Karakesisoglou, I., Hodzic, D. and Olins, D. E. (2009). The LINC-lessgranulocyte nucleus. Eur. J. Cell Biol. 88, 203-214.

    Omary, M. B., Ku, N. O., Tao, G. Z., Toivola, D. M. and Liao, J. (2006). ‘‘Headsand tails’’ of intermediate filament phosphorylation: multiple sites and functionalinsights. Trends Biochem. Sci. 31, 383-394.

    Pagliara, S., Franze, K., McClain, C. R., Wylde, G. W., Fisher, C. L., Franklin,R. J. M., Kabla, A. J., Keyser, U. F. and Chalut, K. J. (2014). Auxetic nuclei inemryonic stem cells. Nat. Mater. (Epub ahead of print) doi:10.1038/nmat3943.

    Pajerowski, J. D., Dahl, K. N., Zhong, F. L., Sammak, P. J. and Discher, D. E.(2007). Physical plasticity of the nucleus in stem cell differentiation. Proc. Natl.Acad. Sci. USA 104, 15619-15624.

    Pittenger, M. F., Mackay, A. M., Beck, S. C., Jaiswal, R. K., Douglas, R., Mosca,J. D., Moorman, M. A., Simonetti, D. W., Craig, S. and Marshak, D. R. (1999).Multilineage potential of adult human mesenchymal stem cells. Science 284,143-147.

    Pugh, G. E., Coates, P. J., Lane, E. B., Raymond, Y. and Quinlan, R. A. (1997).Distinct nuclear assembly pathways for lamins A and C lead to their increaseduring quiescence in Swiss 3T3 cells. J. Cell Sci. 110, 2483-2493.

    Puklin-Faucher, E. and Sheetz, M. P. (2009). The mechanical integrin cycle.J. Cell Sci. 122, 179-186.

    Raab, M., Swift, J., Dingal, P. C. D. P., Shah, P., Shin, J.-W. and Discher, D. E.(2012). Crawling from soft to stiff matrix polarizes the cytoskeleton andphosphoregulates myosin-II heavy chain. J. Cell Biol. 199, 669-683.

    COMMENTARY Journal of Cell Science (2014) 127, 3005–3015 doi:10.1242/jcs.149203

    3014

    http://dx.doi.org/10.1038/ncb2927http://dx.doi.org/10.1038/ncb2927http://dx.doi.org/10.1038/ncb2927http://dx.doi.org/10.1038/ncb2927http://dx.doi.org/10.1016/j.cell.2013.02.031http://dx.doi.org/10.1016/j.cell.2013.02.031http://dx.doi.org/10.1002/term.136http://dx.doi.org/10.1002/term.136http://dx.doi.org/10.1002/term.136http://dx.doi.org/10.1002/cm.20331http://dx.doi.org/10.1002/cm.20331http://dx.doi.org/10.1002/cm.20331http://dx.doi.org/10.1083/jcb.201308029http://dx.doi.org/10.1083/jcb.201308029http://dx.doi.org/10.1083/jcb.201308029http://dx.doi.org/10.1083/jcb.201308029http://dx.doi.org/10.1038/nature10317http://dx.doi.org/10.1038/nature10317http://dx.doi.org/10.1016/0092-8674(90)90470-Yhttp://dx.doi.org/10.1016/0092-8674(90)90470-Yhttp://dx.doi.org/10.1016/j.bpj.2013.09.054http://dx.doi.org/10.1016/j.bpj.2013.09.054http://dx.doi.org/10.1016/j.bpj.2013.09.054http://dx.doi.org/10.1172/JCI38214http://dx.doi.org/10.1172/JCI38214http://dx.doi.org/10.1172/JCI38214http://dx.doi.org/10.1242/jcs.087288http://dx.doi.org/10.1242/jcs.087288http://dx.doi.org/10.1038/nature12105http://dx.doi.org/10.1038/nature12105http://dx.doi.org/10.1038/nature12105http://dx.doi.org/10.1073/pnas.0708572104http://dx.doi.org/10.1073/pnas.0708572104http://dx.doi.org/10.1073/pnas.0708572104http://dx.doi.org/10.1016/j.bpj.2013.01.006http://dx.doi.org/10.1016/j.bpj.2013.01.006http://dx.doi.org/10.1016/j.bpj.2013.01.006http://dx.doi.org/10.1016/j.cub.2013.11.009http://dx.doi.org/10.1016/j.cub.2013.11.009http://dx.doi.org/10.1091/mbc.E11-01-0039http://dx.doi.org/10.1091/mbc.E11-01-0039http://dx.doi.org/10.1091/mbc.E11-01-0039http://dx.doi.org/10.1091/mbc.E11-01-0039http://dx.doi.org/10.1371/journal.pone.0046628http://dx.doi.org/10.1371/journal.pone.0046628http://dx.doi.org/10.1371/journal.pone.0046628http://dx.doi.org/10.1371/journal.pone.0046628http://dx.doi.org/10.4161/nucl.21676http://dx.doi.org/10.4161/nucl.21676http://dx.doi.org/10.4161/nucl.21676http://dx.doi.org/10.4161/nucl.21676http://dx.doi.org/10.1038/ncb0402-e101http://dx.doi.org/10.1038/ncb0402-e101http://dx.doi.org/10.1073/pnas.1111780109http://dx.doi.org/10.1073/pnas.1111780109http://dx.doi.org/10.1073/pnas.1111780109http://dx.doi.org/10.1073/pnas.1111780109http://dx.doi.org/10.1007/s12035-012-8350-1http://dx.doi.org/10.1007/s12035-012-8350-1http://dx.doi.org/10.1007/s12035-012-8350-1http://dx.doi.org/10.1016/j.bpj.2013.04.016http://dx.doi.org/10.1016/j.bpj.2013.04.016http://dx.doi.org/10.1016/j.bpj.2013.04.016http://dx.doi.org/10.1016/j.bpj.2013.04.016http://dx.doi.org/10.1126/science.1211222http://dx.doi.org/10.1126/science.1211222http://dx.doi.org/10.1126/science.1211222http://dx.doi.org/10.1038/cr.2013.118http://dx.doi.org/10.1038/cr.2013.118http://dx.doi.org/10.1016/j.cell.2013.02.028http://dx.doi.org/10.1016/j.cell.2013.02.028http://dx.doi.org/10.1016/j.cell.2013.02.028http://dx.doi.org/10.1016/j.cub.2009.07.069http://dx.doi.org/10.1016/j.cub.2009.07.069http://dx.doi.org/10.1016/j.cub.2009.07.069http://dx.doi.org/10.1016/j.cub.2009.07.069http://dx.doi.org/10.4161/nucl.2.5.17765http://dx.doi.org/10.4161/nucl.2.5.17765http://dx.doi.org/10.4161/nucl.2.5.17765http://dx.doi.org/10.4161/nucl.2.5.17765http://dx.doi.org/10.1016/0022-2836(81)90224-2http://dx.doi.org/10.1016/0022-2836(81)90224-2http://dx.doi.org/10.1016/0022-2836(81)90224-2http://dx.doi.org/10.4161/nucl.2.3.15731http://dx.doi.org/10.4161/nucl.2.3.15731http://dx.doi.org/10.4161/nucl.2.3.15731http://dx.doi.org/10.4161/nucl.2.3.15731http://dx.doi.org/10.4161/nucl.2.3.15731http://dx.doi.org/10.1074/jbc.M513511200http://dx.doi.org/10.1074/jbc.M513511200http://dx.doi.org/10.1074/jbc.M513511200http://dx.doi.org/10.1083/jcb.105.1.577http://dx.doi.org/10.1083/jcb.105.1.577http://dx.doi.org/10.1083/jcb.105.1.577http://dx.doi.org/10.1016/S0006-3495(00)76279-5http://dx.doi.org/10.1016/S0006-3495(00)76279-5http://dx.doi.org/10.1016/0092-8674(92)90245-8http://dx.doi.org/10.1016/0092-8674(92)90245-8http://dx.doi.org/10.1016/0092-8674(92)90245-8http://dx.doi.org/10.1101/gr.159400.113http://dx.doi.org/10.1101/gr.159400.113http://dx.doi.org/10.1101/gr.159400.113http://dx.doi.org/10.1101/gr.159400.113http://dx.doi.org/10.1016/j.cub.2013.10.057http://dx.doi.org/10.1016/j.cub.2013.10.057http://dx.doi.org/10.1016/j.cub.2013.10.057http://dx.doi.org/10.1083/jcb.200804155http://dx.doi.org/10.1083/jcb.200804155http://dx.doi.org/10.1073/pnas.94.3.849http://dx.doi.org/10.1073/pnas.94.3.849http://dx.doi.org/10.1073/pnas.94.3.849http://dx.doi.org/10.1016/j.tibs.2009.03.008http://dx.doi.org/10.1016/j.tibs.2009.03.008http://dx.doi.org/10.1016/S1534-5807(04)00075-9http://dx.doi.org/10.1016/S1534-5807(04)00075-9http://dx.doi.org/10.1016/S1534-5807(04)00075-9http://dx.doi.org/10.1056/NEJMoa0706898http://dx.doi.org/10.1056/NEJMoa0706898http://dx.doi.org/10.1056/NEJMoa0706898http://dx.doi.org/10.1101/gr.141028.112http://dx.doi.org/10.1101/gr.141028.112http://dx.doi.org/10.1101/gr.141028.112http://dx.doi.org/10.1101/gr.141028.112http://dx.doi.org/10.1371/journal.pone.0007050http://dx.doi.org/10.1371/journal.pone.0007050http://dx.doi.org/10.1371/journal.pone.0007050http://dx.doi.org/10.1002/pbc.24788http://dx.doi.org/10.1002/pbc.24788http://dx.doi.org/10.1002/pbc.24788http://dx.doi.org/10.1002/pbc.24788http://dx.doi.org/10.1002/pbc.24788http://dx.doi.org/10.1016/j.bbrc.2004.05.191http://dx.doi.org/10.1016/j.bbrc.2004.05.191http://dx.doi.org/10.1016/j.bbrc.2004.05.191http://dx.doi.org/10.1248/jhs.50.423http://dx.doi.org/10.1248/jhs.50.423http://dx.doi.org/10.1248/jhs.50.423http://dx.doi.org/10.1006/excr.2001.5269http://dx.doi.org/10.1006/excr.2001.5269http://dx.doi.org/10.1006/excr.2001.5269http://dx.doi.org/10.1006/excr.2001.5269http://dx.doi.org/10.1016/j.ejcb.2008.10.001http://dx.doi.org/10.1016/j.ejcb.2008.10.001http://dx.doi.org/10.1016/j.ejcb.2008.10.001http://dx.doi.org/10.1016/j.tibs.2006.05.008http://dx.doi.org/10.1016/j.tibs.2006.05.008http://dx.doi.org/10.1016/j.tibs.2006.05.008http://dx.doi.org/10.1073/pnas.0702576104http://dx.doi.org/10.1073/pnas.0702576104http://dx.doi.org/10.1073/pnas.0702576104http://dx.doi.org/10.1126/science.284.5411.143http://dx.doi.org/10.1126/science.284.5411.143http://dx.doi.org/10.1126/science.284.5411.143http://dx.doi.org/10.1126/science.284.5411.143http://dx.doi.org/10.1242/jcs.042127http://dx.doi.org/10.1242/jcs.042127http://dx.doi.org/10.1083/jcb.201205056http://dx.doi.org/10.1083/jcb.201205056http://dx.doi.org/10.1083/jcb.201205056

  • Jour

    nal o

    f Cel

    l Sci

    ence

    Röber, R. A., Sauter, H., Weber, K. and Osborn, M. (1990). Cells of the cellularimmune and hemopoietic system of the mouse lack lamins A/C: distinctionversus other somatic cells. J. Cell Sci. 95, 587-598.

    Rothballer, A. and Kutay, U. (2013). The diverse functional LINCs of the nuclearenvelope to the cytoskeleton and chromatin. Chromosoma 122, 415-429.

    Rowat, A. C., Jaalouk, D. E., Zwerger, M., Ung,W. L., Eydelnant, I. A., Olins, D. E.,Olins, A. L., Herrmann, H., Weitz, D. A. and Lammerding, J. (2013). Nuclearenvelope composition determines the ability of neutrophil-type cells to passagethrough micron-scale constrictions. J. Biol. Chem. 288, 8610-8618.

    Sawada, Y., Tamada, M., Dubin-Thaler, B. J., Cherniavskaya, O., Sakai, R.,Tanaka, S. and Sheetz, M. P. (2006). Force sensing by mechanical extension ofthe Src family kinase substrate p130Cas. Cell 127, 1015-1026.

    Schwanhäusser, B., Busse, D., Li, N., Dittmar, G., Schuchhardt, J., Wolf, J.,Chen, W. and Selbach, M. (2011). Global quantification of mammalian geneexpression control. Nature 473, 337-342.

    Shackleton, S., Lloyd, D. J., Jackson, S. N. J., Evans, R., Niermeijer, M. F.,Singh, B. M., Schmidt, H., Brabant, G., Kumar, S., Durrington, P. N. et al.(2000). LMNA, encoding lamin A/C, is mutated in partial lipodystrophy. Nat.Genet. 24, 153-156.

    Shimi, T., Pfleghaar, K., Kojima, S., Pack, C. G., Solovei, I., Goldman, A. E.,Adam, S. A., Shumaker, D. K., Kinjo, M., Cremer, T. et al. (2008). The A- andB-type nuclear lamin networks: microdomains involved in chromatinorganization and transcription. Genes Dev. 22, 3409-3421.

    Shin, J.-W., Spinler, K. R., Swift, J., Chasis, J. A., Mohandas, N. and Discher,D. E. (2013). Lamins regulate cell trafficking and lineage maturation of adulthuman hematopoietic cells. Proc. Natl. Acad. Sci. USA 110, 18892-18897.

    Shivashankar, G. V. (2011). Mechanosignaling to the cell nucleus and generegulation. Annual Review of Biophysics 40, 361-378.

    Shoeman, R. L. and Traub, P. (1990). The in vitro DNA-binding properties ofpurified nuclear lamin proteins and vimentin. J. Biol. Chem. 265, 9055-9061.

    Silver, J. T. and Noble, E. G. (2012). Regulation of survival gene hsp70. CellStress Chaperones 17, 1-9.

    Simon, D. N. and Wilson, K. L. (2011). The nucleoskeleton as a genome-associated dynamic ‘network of networks’. Nat. Rev. Mol. Cell Biol. 12, 695-708.

    Simon, D. N., Zastrow, M. S. and Wilson, K. L. (2010). Direct actin binding to A-and B-type lamin tails and actin filament bundling by the lamin A tail. Nucleus 1,264-272.

    Solovei, I., Wang, A. S., Thanisch, K., Schmidt, C. S., Krebs, S., Zwerger, M.,Cohen, T. V., Devys, D., Foisner, R., Peichl, L. et al. (2013). LBR and lamin A/C sequentially tether peripheral heterochromatin and inversely regulatedifferentiation. Cell 152, 584-598.

    Sosa, B. A., Kutay, U. and Schwartz, T. U. (2013). Structural insights into LINCcomplexes. Curr. Opin. Struct. Biol. 23, 285-291.

    Speckman, R. A., Garg, A., Du, F., Bennett, L., Veile, R., Arioglu, E., Taylor, S. I.,Lovett, M. and Bowcock, A. M. (2000). Mutational and haplotype analyses offamilies with familial partial lipodystrophy (Dunnigan variety) reveal recurrentmissense mutations in the globular C-terminal domain of lamin A/C. Am. J. Hum.Genet. 66, 1192-1198.

    Stierlé, V., Couprie, J., Ostlund, C., Krimm, I., Zinn-Justin, S., Hossenlopp, P.,Worman, H. J., Courvalin, J. C. and Duband-Goulet, I. (2003). The carboxyl-terminal region common to lamins A and C contains a DNA binding domain.Biochemistry 42, 4819-4828.

    Sullivan, T., Escalante-Alcalde, D., Bhatt, H., Anver, M., Bhat, N., Nagashima,K., Stewart, C. L. and Burke, B. (1999). Loss of A-type lamin expression

    compromises nuclear envelope integrity leading to muscular dystrophy. J. CellBiol. 147, 913-920.

    Swift, J., Harada, T., Buxboim, A., Shin, J. W., Tang, H. Y., Speicher, D. W. andDischer, D. E. (2013a). Label-free mass spectrometry exploits dozens ofdetected peptides to quantify lamins in wildtype and knockdown cells. Nucleus4, 450-459.

    Swift, J., Ivanovska, I. L., Buxboim, A., Harada, T., Dingal, P. C. D. P., Pinter, J.,Pajerowski, J. D., Spinler, K. R., Shin, J.-W., Tewari, M. et al. (2013b).Nuclear lamin-A scales with tissue stiffness and enhances matrix-directeddifferentiation. Science 341, 1240104.

    Talwar, S., Kumar, A., Rao, M., Menon, G. I. and Shivashankar, G. V. (2013).Correlated spatio-temporal fluctuations in chromatin compaction statescharacterize stem cells. Biophys. J. 104, 553-564.

    Tilli, C. M., Ramaekers, F. C. S., Broers, J. L. V., Hutchison, C. J. and Neumann,H. A. M. (2003). Lamin expression in normal human skin, actinic keratosis,squamous cell carcinoma and basal cell carcinoma.Br. J. Dermatol. 148, 102-109.

    Toivola, D. M., Strnad, P., Habtezion, A. and Omary, M. B. (2010). Intermediatefilaments take the heat as stress proteins. Trends Cell Biol. 20, 79-91.

    Viano, D. C., Casson, I. R., Pellman, E. J., Zhang, L., King, A. I. and Yang, K. H.(2005). Concussion in professional football: brain responses by finite elementanalysis: part 9. Neurosurgery 57, 891-916, discussion 891-916.

    Wagner, N. and Krohne, G. (2007). LEM-Domain proteins: new insights intolamin-interacting proteins. Int. Rev. Cytol. 261, 1-46.

    Wang, H. B., Dembo, M. and Wang, Y. L. (2000). Substrate flexibility regulatesgrowth and apoptosis of normal but not transformed cells. Am. J. Physiol. 279,C1345-C1350.

    Watt, F. M. and Huck, W. T. S. (2013). Role of the extracellular matrix in regulatingstem cell fate. Nat. Rev. Mol. Cell Biol. 14, 467-473.

    Wazir, U., Ahmed, M. H., Bridger, J. M., Harvey, A., Jiang, W. G., Sharma, A. K.and Mokbel, K. (2013). The clinicopathological significance of lamin A/C, laminB1 and lamin B receptor mRNA expression in human breast cancer. Cell. Mol.Biol. Lett. 18, 595-611.

    Wilson, K. L. and Berk, J. M. (2010). The nuclear envelope at a glance. J. CellSci. 123, 1973-1978.

    Wilson, K. L. and Foisner, R. (2010). Lamin-binding Proteins. Cold Spring Harb.Perspect. Biol. 2, a000554.

    Winer, J. P., Janmey, P. A., McCormick, M. E. and Funaki, M. (2009). Bonemarrow-derived human mesenchymal stem cells become quiescent on softsubstrates but remain responsive to chemical or mechanical stimuli. Tissue Eng.Part A 15, 147-154.

    Wolf, K., Te Lindert, M., Krause, M., Alexander, S., Te Riet, J., Willis, A. L.,Hoffman, R. M., Figdor, C. G., Weiss, S. J. and Friedl, P. (2013). Physicallimits of cell migration: control by ECM space and nuclear deformation andtuning by proteolysis and traction force. J. Cell Biol. 201, 1069-1084.

    Worman, H. J. (2012). Nuclear lamins and laminopathies. J. Pathol. 226, 316-325.Xu, R., Boudreau, A. and Bissell, M. J. (2009). Tissue architecture and function:dynamic reciprocity via extra- and intra-cellular matrices. Cancer MetastasisRev. 28, 167-176.

    Zuela, N., Bar, D. Z. and Gruenbaum, Y. (2012). Lamins in development, tissuemaintenance and stress. EMBO Rep. 13, 1070-1078.

    Zullo, J. M., Demarco, I. A., Piqué-Regi, R., Gaffney, D. J., Epstein, C. B.,Spooner, C. J., Luperchio, T. R., Bernstein, B. E., Pritchard, J. K., Reddy, K. L.et al. (2012). DNA sequence-dependent compartmentalization and silencing ofchromatin at the nuclear lamina. Cell 149, 1474-1487.

    COMMENTARY Journal of Cell Science (2014) 127, 3005–3015 doi:10.1242/jcs.149203

    3015

    http://dx.doi.org/10.1007/s00412-013-0417-xhttp://dx.doi.org/10.1007/s00412-013-0417-xhttp://dx.doi.org/10.1074/jbc.M112.441535http://dx.doi.org