<<

Cecilia Bergqvist The genetic material is highly structured within the nucleus, with transcriptionally inactive heterochromatin enriched at the nuclear The role of periphery and active euchromatin in the nuclear interior. The together with several hundred nuclear envelope transmembrane in proteins (NETs) connect chromatin to the nuclear periphery. Most NETs are tissue-specific and uncharacterized, with mutations linked to organization, differentiation and The role of nuclear envelope proteins in chromatin organization, differentiation and disease distinct degenerative disorders, referred to as . The NET disease primarily studied in this thesis is called Spindle-Associated Membrane 1 (Samp1). We showed that overexpression of Samp1 induced a fast differentiation of human induced pluripotent stem cells and that the binding between two NETs, Samp1 and , is regulated by Cecilia Bergqvist RanGTP. Another focus of this thesis was the development of a novel method, Fluorescent Ratiometric Imaging of Chromatin (FRIC). FRIC quantitatively monitors the epigenetic state of chromatin in live cells. Using FRIC, we were able to show that Samp1 promotes peripheral heterochromatin organization. FRIC also detected an increased distribution of heterochromatin at the nuclear periphery during neuronal differentiation. In conclusion, FRIC is a useful tool that could serve medical research in elucidating the effects of different chemical agents and NE proteins in chromatin organization.

Cecilia Bergqvist

ISBN 978-91-7911-230-1

Department of and Biophysics

Doctoral Thesis in Neurochemistry with Molecular Neurobiology at Stockholm University, Sweden 2020

The role of nuclear envelope proteins in chromatin organization, differentiation and disease Cecilia Bergqvist Academic dissertation for the Degree of Doctor of Philosophy in Neurochemistry with Molecular Neurobiology at Stockholm University to be publicly defended on Friday 2 October 2020 at 10.00 in Magnélisalen, Kemiska övningslaboratoriet, Svante Arrhenius väg 16 B.

Abstract In the genetic material is separated from the by the nuclear envelope (NE), consisting of the outer and inner nuclear membrane, the nuclear lamina and the nuclear pores. The genetic material is highly structured with transcriptionally inactive heterochromatin enriched at the nuclear periphery and transcriptionally active euchromatin in the nuclear interior. Underlying the inner nuclear membrane is the nuclear lamina (nucleoskeleton) that together with several hundred nuclear envelope transmembrane proteins (NETs) connect chromatin to the nuclear periphery. Most NETs are uncharacterized and expressed in a tissue-specific manner. Mutations in NE proteins are linked to distinct degenerative disorders, referred to as envelopathies or laminopathies. The NET primarily studied in this thesis is called Spindle-Associated 1 (Samp1). We showed that overexpression of Samp1 induced a fast differentiation of human induced pluripotent stem cells and that the binding between two NETs, Samp1 and Emerin, is regulated by RanGTP. Another focus of this thesis was the development and use of a novel method called Fluorescent Ratiometric Imaging of Chromatin (FRIC). FRIC quantitatively monitors the epigenetic state of chromatin in live cells. Using FRIC, we were able to show that Samp1 promotes peripheral heterochromatin organization. FRIC also detected an increased distribution of heterochromatin at the nuclear periphery during neuronal differentiation. In conclusion, FRIC is a useful tool that could serve medical research in elucidating the effects of different chemical agents and the roles of NE proteins in chromatin organization.

Keywords: Nuclear envelope proteins, chromatin organization, epigenetics, differentiation, quantitative image analysis, Samp1.

Stockholm 2020 http://urn.kb.se/resolve?urn=urn:nbn:se:su:diva-184182

ISBN 978-91-7911-230-1 ISBN 978-91-7911-231-8

Department of Biochemistry and Biophysics

Stockholm University, 106 91 Stockholm

THE ROLE OF NUCLEAR ENVELOPE PROTEINS IN CHROMATIN ORGANIZATION, DIFFERENTIATION AND DISEASE

Cecilia Bergqvist

The role of nuclear envelope proteins in chromatin organization, differentiation and disease

Cecilia Bergqvist ©Cecilia Bergqvist, Stockholm University 2020

ISBN print 978-91-7911-230-1 ISBN PDF 978-91-7911-231-8

Cover image: A colony of human induced pluripotent stem cells stained for Emerin (cyan) and A/C (red).

Printed in Sweden by Universitetsservice US-AB, Stockholm 2020 Distributor: Department of Biochemistry and Biophysics, Stockholm University Grandma, perhaps the stars are openings in the sky where our loved ones shine to let us know they are there.

Abstract

In eukaryotes the genetic material is separated from the cytoplasm by the nuclear envelope (NE), consisting of the outer and inner nuclear membrane, the nuclear lamina and the nuclear pores. The genetic material is highly structured with transcriptionally inactive heterochromatin enriched at the nuclear periphery and transcriptionally active euchromatin in the nuclear interior. Underlying the inner nuclear membrane is the nuclear lamina (nucleoskeleton) that together with several hundred nuclear envelope transmembrane proteins (NETs) connect chromatin to the nuclear periphery. Most NETs are uncharacterized and expressed in a tissue-specific manner. Mutations in NE proteins are linked to distinct degenerative disorders, referred to as envelopathies or laminopathies. The NET primarily studied in this thesis is called Spindle-Associated Membrane Protein 1 (Samp1). We showed that overexpression of Samp1 induced a fast differentiation of human induced pluripotent stem cells and that the binding between two NETs, Samp1 and Emerin, is regulated by RanGTP. Another focus of this thesis was the development and use of a novel method called Fluorescent Ratiometric Imaging of Chromatin (FRIC). FRIC quantitatively monitors the epigenetic state of chromatin in live cells. Using FRIC, we were able to show that Samp1 promotes peripheral heterochromatin organization. FRIC also detected an increased distribution of heterochromatin at the nuclear periphery during neuronal differentiation. In conclusion, FRIC is a useful tool that could serve medical research in elucidating the effects of different chemical agents and the roles of NE proteins in chromatin organization.

1

2

List of publications

This thesis is based on the following publications and manuscripts, referred to as Paper I-V in the text:

Paper I: Cecilia Bergqvist, Frida Niss, Ricardo A Figueroa, Marie Beckman, Danuta Maksel, Mohammed H Jafferali, Agné Kulyté, Anna-Lena Ström, and Einar Hallberg (2019). Monitoring of Chromatin Organization in Live Cells by FRIC. Effects of the Inner Nuclear Membrane Protein Samp1. Nucleic Acids Research 47, no. 9: e49. https://doi.org/10.1093/nar/gkz123.

Paper II: Balaje Vijayaraghavan, Ricardo A. Figueroa, Cecilia Bergqvist, Amit J. Gupta, Paulo Sousa, and Einar Hallberg (2018). RanGTPase Regulates the Interaction between the Inner Nuclear Membrane Proteins, Samp1, and Emerin. Biochimica et Biophysica Acta (BBA) - Biomembranes 1860, no. 6: 1326–34. https://doi.org/10.1016/j.bbamem.2018.03.001.

Paper III: Cecilia Bergqvist, Mohammed Hakim Jafferali, Santhosh Gudise, Robert Markus, and Einar Hallberg (2017). An Inner Nuclear Membrane Protein Induces Rapid Differentiation of Human Induced Pluripotent Stem Cells. Stem Research 23: 33–38. https://doi.org/10.1016/j.scr.2017.06.008.

Paper IV: Cecilia Bergqvist, Urška Kašnik, and Einar Hallberg. Chromatin reorganization during neuronal differentiation. Manuscript.

Paper V: Cecilia Bergqvist, Urška Kašnik, and Einar Hallberg. Investigations of Emery-Dreifuss Muscular Dystrophy mutants of Samp1. Manuscript.

Minor parts of the work presented in this thesis have previously been published in my licentiate thesis: Cecilia Bergqvist, The role of nuclear membrane proteins in differentiation and chromatin organization (2016), Stockholm University. ISBN 978-91-7649-632-9.

3

Table of Contents

Abstract ...... 1 List of publications ...... 3 Abbreviations ...... 7 1.1 The nuclear envelope ...... 9 1.1.1 The nuclear lamina ...... 10 Lamin A processing ...... 10 The function of the nuclear lamina ...... 11 1.1.2 The linker of nucleoskeleton and (LINC) complex ...... 12 The function of the LINC complex ...... 13 1.1.3 Nuclear envelope transmembrane proteins ...... 13 Emerin ...... 13 INM proteins in myogenesis ...... 14 Nucleocytoplasmic transport of proteins to the inner nuclear membrane (INM) ...... 14 1.1.4 RanGTPase in nucleocytoplasmic transport and ...... 15 1.1.5 Chromatin ...... 16 variants and post-translational modifications ...... 17 Tethering chromatin to the nuclear periphery ...... 17 Chromatin organization during differentiation ...... 18 1.2 Diseases of the nuclear envelope - Emery-Dreifuss Muscular Dystrophy ...... 19 1.3 The inner nuclear membrane protein Samp1 ...... 21 Samp1 in Emery-Dreifuss Muscular Dystrophy ...... 22 Samp1 and RanGTPase...... 22 Tissue-specific expression of Samp1 ...... 22 2. Aim ...... 23 3. Methodological considerations ...... 24 3.1 Cell types ...... 24 3.2 Live cell imaging ...... 24 3.3 Fluorescent Ratiometric Imaging of Chromatin (FRIC) ...... 25 3.4 Fluorescence recovery after photobleaching (FRAP) ...... 25 3.5 knock out using CRISPR/Cas9 mediated editing ...... 26 3.6 Microscale thermophoresis (MST) ...... 26 3.7 Super-resolution using Structured Illumination Microscopy (SIM) ...... 26

4

4. Results and Discussion ...... 27 Monitoring chromatin organization at the nuclear periphery (Paper I) ...... 27 The effect of Samp1 on chromatin distribution (Paper I, IV and V) ...... 28 Interactions between nuclear envelope proteins (Paper II) ...... 28 Samp1 in cell differentiation (Paper III and IV) ...... 29 Monitoring chromatin organization during differentiation (paper IV) ...... 30 Emery-Dreifuss Muscular Dystrophy mutations of Samp1 (paper V) ...... 30 5. Conclusions ...... 32 6. Future Investigations ...... 33 7. Populärvetenskaplig sammanfattning ...... 35 8. Acknowledgements ...... 36 9. References ...... 38

5

6

Abbreviations aa amino acids AA anacardic acid AD-EDMD autosomal dominant EDMD BAF barrier-to-autointegration factor CRISPR/Cas clustered, regularly interspaced, short, palindromic repeats/CRISPR associated system CT. chaetomium thermophilum CNS central nervous system DamID DNA adenine identification DNA deoxyribonucleic acid dsDNA double-stranded deoxyribonucleic acid EDMD Emery-Dreifuss muscular dystrophy ER FP fluorescent protein FRAP fluorescence recovery after photobleaching FRIC fluorescence ratiometric imaging of chromatin GDP/GTP guanosine diphosphate/guanosine triphosphate GFP/YFP/EGFP green/yellow/enhanced green fluorescent protein HAT histone acetyltransferase HDAC histone deacetylase HGPS Hutchinson–Gilford syndrome hiPSC human induced pluripotent stem cell HP1 heterochromatin protein 1 ID intrinsic disorder IF immunofluorescence IP immunoprecipitation KASH Klarsicht, ANC1, Syne1 homology INM inner nuclear membrane LAD lamina-associated domain LBR Lamin B receptor LEM Lap2, Emerin, Man1 LINC linker of nucleoskeleton and cytoskeleton MCLIP membrane protein crosslink immunoprecipitation MST microscale thermophoresis NE nuclear envelope NET nuclear envelope NES nuclear export signal NGF nerve growth factor NLS nuclear localization signal NPC complex NTR nuclear transport receptor Nups ONM outer nuclear membrane PNS perinuclear space RA retinoic acid RanGAP1 RanGTPase-activating protein 1 RCC1 regulator of chromosome condensation 1 Samp1 spindle associated membrane protein 1 SIM structured illumination microscopy TAN lines transmembrane actin-associated nuclear lines LAP lamina-associated polypeptide S.pombe schizosaccharomyces pombe SAF spindle assembly factor SAHF senescence-associated heterochromatin foci sgRNA single guided ribonucleic acid SUN Sad1/UNC-84 TSA trichostatin A X-EDMD recessive X-linked EDMD

7

8

1.1 The nuclear envelope

The largest in eukaryotic cells is the nucleus, which contains most of the cell’s genetic material. The nucleus is surrounded by the nuclear envelope (NE), a highly complex structure that separates the from the cytoplasm (Burke and Stewart, 2002; Hetzer, 2010; Kite, 1913; Stewart et al., 2007). The NE has two concentric membranes, the outer nuclear membrane (ONM) that is an extension of the rough endoplasmic reticulum and the inner nuclear membrane (INM) that connects the nuclear lamina and chromatin to the nuclear periphery (Amendola and van Steensel, 2014; Solovei et al., 2013). Transport of proteins and RNA across the NE (nucleocytoplasmic transport) occurs solely through the macromolecular structures called nuclear pore complexes (NPCs) (Ibarra and Hetzer, 2015; Jahed et al., 2016). Historically, the NE was viewed as a mere physical container of the genomic material (Burke and Stewart, 2014). However, with an increasing number of human diseases linked to NE proteins, the NE is now recognized to perform many distinct and important functions (Burke and Stewart, 2002; Dauer and Worman, 2009, 2009; Vlcek and Foisner, 2007; Worman and Schirmer, 2015). Several hundred unique integral membrane proteins have been categorized as potential NE proteins (Korfali et al., 2010; Schirmer et al., 2003) and many of them display tissue- specific expression patterns (Korfali et al., 2012; Wilkie et al., 2011). The few NE proteins studied, are linked to functions involving chromatin organization, , proliferation, differentiation, cytoskeleton organization, mechanical stability, and cell migration (Hetzer, 2010; Parada et al., 2004; Taddei et al., 2004; Wilson and Berk, 2010).

The INM protein primarily studied in this thesis is called Spindle-Associated Membrane Protein 1 (Samp1) (Buch et al., 2009), figure 1. Samp1 is associated with the nuclear lamina (the nucleoskeleton), a meshwork of intermediate filaments that gives the nucleus structure and stability (see 1.1.1 The nuclear lamina) (Buch et al., 2009; Gudise et al., 2011). Samp1 also interacts with the linker of nucleoskeleton and cytoskeleton (LINC) complex, which is important in nuclear migration and mechanical signaling (Borrego-Pinto et al., 2012; Gudise et al., 2011) (see 1.1.2 The LINC complex). Samp1 binds to the INM protein Emerin (see 1.1.3 Nuclear envelope transmembrane proteins) (Jafferali et al., 2014; paper II) , linked to Emery-Dreifuss muscular dystrophy (EDMD) (Bione et al., 1994) (see 1.2 Diseases of the nuclear envelope). Samp1 is mislocalized (Mattioli et al., 2018) and mutated in EDMD patients (Meinke et al., 2020). Samp1 additionally binds to the nuclear G-protein, RanGTPase, which plays fundamental roles in nucleocytoplasmic transport, assembly of the mitotic spindle, and postmitotic nuclear reformation (Melchior, 2001; Vijayaraghavan et al., 2016) (see 1.1.4 RanGTPase in nucleocytoplasmic transport and mitosis).

9

Figure 1. Samp1 binds directly to the inner nuclear membrane protein Emerin and the nuclear G-protein . Samp1 also interacts with the nuclear lamina and the LINC complex (Sun1 and Sun2) (Cecilia Bergqvist, 2020, BioRender).

1.1.1 The nuclear lamina Underlying the nuclear envelope is the nuclear lamina that forms a complex meshwork of intermediate filaments that gives the nucleus its structure and stability (Burke and Stewart, 2014; Dechat et al., 2008). These filaments are composed of encoded by LMNA (splice variants Lamin A and C), LMNB1 (), and LMNB2 (Lamin B2), which can form separate structural networks (Dechat et al., 2008; Guerreiro and Kind, 2019; Nmezi et al., 2019). B-type lamins are essential proteins expressed in all mammalian tissues (Broers et al., 1997), while A-type lamins are undetectable by immunofluorescence in embryonic cells and often used as markers for differentiation (Machiels et al., 1996; Zhang et al., 2011).

Lamin A processing Lamin A is post-translationally modified from pre-Lamin A. Pre-Lamin A is first farnesylated at the cysteine residue in the C-terminal -CSIM sequence. Then the -SIM sequence is cleaved off and the carboxylic acid group (COOH) of the cysteine is methylated. Pre-Lamin A is further processed by Zmpste24, which removes the last 15 amino acids containing the farnesyl-group, resulting in mature Lamin A (De Sandre-Giovannoli et al., 2003; Musich and Zou, 2009), see figure 2 below. The farnesyl-group associates pre-Lamin A to the INM (Ho and Hegele, 2019).

10

Figure 2. Post-translational modifications of pre-lamin A, were pre-lamin A is processed to form mature Lamin A or the incompletely processed, disease-causing, form (Bergqvist, 2016).

Hutchinson-Gilford progeria syndrome (HGPS) is a disorder where progerin, a truncated incompletely processed form of Lamin A (Eriksson et al., 2003) see figure 2, accumulates at the INM (De Sandre-Giovannoli et al., 2003; Merideth et al., 2008; Musich and Zou, 2009). HGPS is a rare, fatal, sporadic, autosomal dominant disease with early onset of symptoms resembling premature aging. Most patients (90%) have a silent mutation (p.G608G), which leads to a cryptic splice site resulting in a truncated pre-Lamin A form called progerin that lacks 50 internal amino acids (Eriksson et al., 2003; Merideth et al., 2008) and consequently the Zmpste24 cleavage site. This results in the accumulation of farnesylated progerin at the INM, which disrupts the nuclear lamina and alters transcription (Ho and Hegele, 2019). Cells that accumulate progerin are hypersensitive to mechanical stress, a typical phenotype in lamin dysfunction (Gruenbaum and Foisner, 2015). A nonfunctional Lamin A network also leads to disruption of chromatin tethering to the nuclear envelope (van Steensel and Belmont, 2017) which leads to loss of heterochromatin (see 1.1.5 Chromatin) at the nuclear periphery (Lattanzi et al., 2014).

The function of the nuclear lamina Lamins maintain the shape and mechanical stability of the nucleus (Amendola and van Steensel, 2014). Lamin A/C deficient cells display both impaired mechanotransduction (Houben et al., 2007) and cell migration (Chen et al., 2018). This is explained mainly by the destabilization of the LINC complex (see 1.1.2 The LINC complex). Cells lacking Lamin A/C are more deformable than cells expressing Lamin A/C (Hah and Kim, 2019). Leukocytes that lack A-type lamins can squeeze through narrow constrictions between cells and the extracellular matrix (Harada et al., 2014). Additionally, mammalian erythrocytes lack nuclei completely, which facilitate deformation when squeezing through narrow capillaries (Ji et al., 2011). Lamin B1 and Lamin B2 are essential for brain development. Mice deficient in Lamin B1, have abnormal neuronal migration and reduced numbers of neurons while Lamin B2 deficiency leads to nuclear elongation (Coffinier et al., 2011). Furthermore, the depletion of Lamin B1 causes the nucleus to spin around within the cell which suggests impairment of nuclear anchoring (see 1.1.2 LINC complex). These cells also show impaired proliferation and chromatin organization (Hah and Kim, 2019).

Around 40% of the genome interacts with nuclear lamins (Peric-Hupkes et al., 2010) where lamins generally tether chromatin to the nuclear periphery. This usually occurs in a 11 transcriptionally silencing way through a Lamin A/C (A-tether) and Lamin B receptor (LBR) (B-tether) dependent manner (Solovei et al., 2013). In mammalian cells, there are around 1000- 1500 lamina-associated domains (LADs) (van Steensel and Belmont, 2017), regions of the genome connected to lamins. Most LADs are gene-poor or poorly transcribed and have histone marks of a repressed chromatin state (Guerreiro and Kind, 2019). During differentiation in general, chromatin becomes increasingly tethered to the nuclear periphery (Talwar et al., 2013). In contrast to Lamin A/C, the LBR has been shown to prevent differentiation (Sola Carvajal et al., 2015; Solovei et al., 2013) and is for example downregulated in neuronal development (Clowney et al., 2012). As LBR expression decreases with differentiation, Lamin A/C expression increases (Nikolakaki et al., 2017; Sola Carvajal et al., 2015; Solovei et al., 2013). Myoblast transcriptome analyses from mice depleted of LBR or Lamin A/C showed that the LBR- and lamin-A-dependent heterochromatin tethers had the opposite effect on muscle-specific gene expression. Loss of LBR increased expression while the loss of Lamin A/C decreased expression (Nikolakaki et al., 2017; Solovei et al., 2013). One example of how tethering to the nuclear periphery can prevent premature differentiation is MyoD, one of the earliest markers of myogenic commitment. First, during myogenesis, the MyoD locus relocates from the nuclear periphery to the nuclear interior where it is transcribed and promotes terminal differentiation of the myogenic cells (Yao et al., 2011). The LBR directly binds B-type lamins and chromatin (Nikolakaki et al., 2017), whereas Lamin A/C indirectly tethers chromatin through chromatin-binding proteins such as the INM proteins (Amendola and van Steensel, 2014; Solovei et al., 2013; Thanisch et al., 2017). Seven INM proteins (Lap2β, Man1, LEMD2, Emerin, Lap1β, Lap2α, and Samp1) were tested for being the chromatin binding component in A-tethering. Only LEMD2 expression correlated with Lamin A/C expression in various tissues, for example, the rod cells of nocturnal (Thanisch et al., 2017). These cells lack both A-tethers and B-tethers, which results in an inverted chromatin state (with heterochromatin concentrated in the nuclear interior in so-called chromocenters that acts as lenses), which improves the light transmission of the photoreceptors in the retina and hence night vision (Solovei et al., 2013). LEMD2 was dependent of Lamin A/C for localization to the NE (Thanisch et al., 2017) and directly binds chromatin (Barrales et al., 2016). However, expression of both LEMD2 and A-type lamins did not result in restored peripheral heterochromatin in the rod cells of the nocturnal animals (Thanisch et al., 2017). This suggests that A-tethering might be more complex and include more than one chromatin binding component. Considering the number of binding partners for Lamin A/C and its role in gene regulation during differentiation, the chromatin binding components of the A-tether probably differ between tissues.

1.1.2 The linker of nucleoskeleton and cytoskeleton (LINC) complex The linker of nucleoskeleton and cytoskeleton (LINC) complex provides a direct physical link between the cytoskeleton and the nucleoskeleton (nuclear lamina). This enables mechanical signaling to be transferred directly across the NE (Crisp et al., 2006). The LINC complex consists of nesprins that are Klarsicht, ANC1, and Syne homology (KASH) domain proteins that span the ONM and Sad and UNC-84 (SUN) domain proteins that span the INM (Chang et al., 2015a; Crisp et al., 2006). Trimers of the SUN domain proteins and KASH domain proteins bind in the perinuclear space (PNS), between the ONM and INM. Together with the NPCs, the LINC complex maintains the PNS thickness to a constant 30-50 nm (Jahed et al., 2016). The nucleoplasmic domain of the SUN domain proteins binds directly to both lamins, chromatin, and some of the NETs (Burke and Stewart, 2014). In mammals, there are five known SUN domain proteins and six known KASH domain proteins. KASH domain proteins interact with cytoskeleton components such as actin, intermediate filaments and microtubules at the ONM (Jahed et al., 2016). The presence of different KASH-domain

12 proteins and SUN-domain protein isoforms allows for tissue-specificity and different functional roles together with their interaction partners at the INM and ONM (Chang et al., 2015b).

The function of the LINC complex The LINC complex is essential in nuclear positioning during cell migration and nuclear migration, occurring both in for example neuronal and muscle differentiation (Bouzid et al., 2019; Burke, 2019). In mouse NIH/3T3 , the INM protein Samp1 (see 1.3 The inner nuclear membrane protein Samp1) is a part of the transmembrane actin-associated nuclear (TAN) lines (Borrego-Pinto et al., 2012). These TAN lines consist of LINC complexes that are connected to the nuclear lamina and actin filaments and are aligned perpendicularly against the leading edge in migrating cells. TAN lines are essential for positioning the in front of the nucleus, at the leading edge, before cell migration and movement of the nucleus (Borrego-Pinto et al., 2012). Samp1 depletion results in a detachment of the centrosome from the nuclear periphery (Buch et al., 2009), a common phenotype in Emery-Dreifuss Muscular Dystrophy (EDMD) patient cells and cells depleted of functional Lamin A/C or Emerin or SUN1 and SUN2 or Nesprin1 and Nesprin2 (Chang et al., 2015a, 2015b; Salpingidou et al., 2007). This indicates that Samp1 and these proteins work together in positioning the centrosome close to the nuclear surface (Gudise et al., 2011). Samp1 is a component of the TAN lines consisting of SUN2, Nesprin2G, and Lamin A/C. The depletion of Samp1 destabilizes the LINC complexes in TAN lines which impairs nuclear movement and therefore cell migration in wound-healing assays (Borrego-Pinto et al., 2012). In , depletion of Ima1 (homolog to Samp1 in Schizosaccharomyces pombe) has a similar function as it disrupts the Sad1-Kms2 complex (a LINC complex homolog in yeast). This results in inefficient tethering of centromeric heterochromatin to the nuclear periphery and less tolerance to mechanical stress (Steglich et al., 2012). Nuclear movement is also an important process in developmental and cellular processes, such as muscle and neuronal differentiation (Crisp et al., 2006; Dechat et al., 2008). For example, depletion of SUN1 and SUN2 results in deficient neuronal migration which leads to prenatal death in rodents due to deficient neuronal development of the CNS (Crisp et al., 2006).

1.1.3 Nuclear envelope transmembrane proteins Nuclear envelope transmembrane proteins (NETs) are proteins that are embedded in the nuclear membrane. Around 10% of the transmembrane proteins in eukaryotes are nuclear (Mudumbi et al., 2020). There are several hundreds of different NETs, some are ubiquitously expressed while others are expressed in a tissue-specific pattern (Korfali et al., 2012; Zuleger et al., 2013). In general, the nuclear envelope proteome is highly tissue specific. Comparing three tissues (muscle, liver, and leukocytes) only 16% of the NETs were shared (Korfali et al., 2012). Diseases connected to mutations in NETs usually have a tissue-specific pathology, for example, some mutations in ubiquitously expressed Man1 increase bone density, and duplication of the Lamin B1 locus results in demyelination of the CNS (Janin et al., 2017; Worman and Schirmer, 2015). Only a few of the NETs have been characterized. Out of these, most were found to bind to the nuclear lamina, chromatin, or chromatin-binding proteins (Amendola and van Steensel, 2014; Harr et al., 2016; Vlcek and Foisner, 2007).

Emerin Emerin (one of the direct binding partners of Samp1) interacts with chromatin through chromatin-binding proteins such as barrier-to-autointegration factor (BAF) (Margalit et al., 2007; Samson et al., 2017) and histone deacetylase 3 (HDAC3) (Demmerle et al., 2012). These 13 proteins together with lamins repress transcription at the nuclear periphery (Berk et al., 2013). Emerin is a Lap2, Emerin, Man1 (LEM) domain protein, with a conserved helix-loop-helix domain consisting of 40 residues (Berk et al., 2013). LEM-domain proteins are known to compensate for each other, for example in X-EDMD patients (depleted of Emerin) Lap2 is upregulated (Koch and Holaska, 2014). Emerin is dependent on Lamin A/C and Samp1 for localization to the NE (Buch et al., 2009; Gudise et al., 2011; paper II) and binds SUN-domain proteins (Berk et al., 2013).

INM proteins in myogenesis Mice depleted of both Lamin A and C, die 8 weeks after birth (Sullivan et al., 1999), whereas mice depleted of only Lamin A and consequently pre-Lamin A are entirely healthy (Fong et al., 2006). Although not essential in mice, pre-Lamin A has been shown to have a function in myogenesis (Mattioli et al., 2011). Myogenesis (formation of muscular tissue) proceeds in different stages. First, myoblasts exit the and start to express muscle- specific transcription factors and myogenic structural proteins. Then these “committed” myoblasts migrate to align longitudinally next to each other where their plasma membranes fuse to form mature myotubules (Chal and Pourquié, 2017; Fiorotto, 2012). The nuclear movement during this fusion depends on the NE composition at the nuclear poles which are adjacent to neighboring nuclei (Mattioli et al., 2011). Farnesylated pre-Lamin A is concentrated at nuclear poles of committed myoblasts and myotubes, where it concentrates SUN2 to the nuclear poles. In the absence of farnesylated pre-Lamin A, nuclear positioning in the formed myotubes is affected which resulted in nuclear clustering and misshaping (Mattioli et al., 2011). In cells and patients with EDMD, with reduced levels of pre-Lamin A, Samp1 enrichment is absent from the nuclear poles, suggesting Samp1 and pre-Lamin A might function together in myogenesis (Mattioli et al., 2018). Samp1 displays tissue-specific expression with high expression in the brain, muscle, and testis (Korfali et al., 2012; Thanisch et al., 2017; Zuleger et al., 2013). During myogenesis, Samp1a expression increases seven-fold and knockdown of all Samp1 splice variants (Samp1a, Samp1b and Samp1c) (see 1.3 Samp1), completely blocks myogenesis, an effect that could be rescued by ectopic expression of Samp1a alone (Jafferali et al., 2017). In comparison, depletion of some LEM-domain proteins (Lem2, Emerin, and Man1, whose expression levels also increase during myogenesis) only caused 2-4-fold inhibition of myogenesis (Huber et al., 2009). Samp1 depleted cells lack expression of myogenin and myogenic structural differentiation marker myosin heavy chain (MyHC). These cells were also impaired in the ability to exit the cell cycle. Similar effects were observed after the depletion of the LEM- domain proteins or Lamin A/C (Huber et al., 2009; Muchir et al., 2009b, 2009a). In all cases, this was associated with the hyperactivation of MAPK kinase signaling, which was also the case in Samp1 depleted myoblasts (Jafferali et al., 2017). Other muscle-specific NETs such as the NET39, Tmem38a, and WFS1 have also been implicated in myogenesis. Together these three NETs affect 37% of all genes that change during myogenesis and combined knock-down results in an almost blocked myogenesis (Robson et al., 2016), suggesting that several NETs collaborate in tissue-specific development.

Nucleocytoplasmic transport of proteins to the inner nuclear membrane (INM) Most INM proteins are synthesized and co-translationally inserted into the RER membrane, which is continuous with the ONM (Ibarra and Hetzer, 2015). Nucleocytoplasmic transport of soluble proteins is well-studied and occurs through the central channel of the nuclear pore complex (NPC) (Fahrenkrog and Aebi, 2003). However, various models on how INM-proteins are transported to the INM have been proposed (Katta et al., 2014). These INM proteins can be transported into the nucleus either via the central or peripheral channels of the

14

NPC. The NPC is a highly conserved megadalton (125 MDa) macromolecular structure composed of multiple copies of 30 different proteins called nucleoporins (Nups), together forming an eight-fold symmetry with one large central channel and 8 smaller peripheral channels (Ibarra and Hetzer, 2015). Most of the INM-proteins use the peripheral channels. However, some of them contain nuclear localization signals (NLS) and an intrinsic disorder (ID) region which enables transport through the central channel (see 1.1.4 RanGTPase in nucleocytoplasmic transport and mitosis) (Mudumbi et al., 2020). Mutating the NLS of these proteins or blocking the central channel, results in a slower transport through the peripheral channels instead (Mudumbi et al., 2020). Blocking the peripheral channels, however, blocks nucleocytoplasmic transport of INM- proteins completely (Mudumbi et al., 2020). In the free lateral diffusion-retention model, INM-proteins with a nucleoplasmic domain smaller than 60 kDa (such as LBR 22kDa (Nikolakaki et al., 2017), Emerin 26 kDa (Berk et al., 2013) or Samp1 26kDa) can freely pass the small (around 10 nm wide) peripheral channels. INM-proteins stay embedded in the nuclear membrane through the translocation across the NPC and are retained in the INM by binding to chromatin, other INM-proteins, or the nuclear lamina (Katta et al., 2014). Only around 9% of the hundreds of INM-proteins could potentially use the NLS-dependent facilitated transport model instead (Mudumbi et al., 2020). These proteins have an NLS and a flexible structure (ID region) that could stretch through the peripheral channels to the central channel where the NLS is bound to nuclear transport receptors (NTRs) such as importins. NTRs facilitate transport through the larger central channel (around 40-50 nm wide) via a mechanism depending on RanGTP (Katta et al., 2014; Maimon et al., 2012; Mudumbi et al., 2020) (see 1.1.4 RanGTPase in nucleocytoplasmic transport and mitosis).

1.1.4 RanGTPase in nucleocytoplasmic transport and mitosis The small GTPase called Ras-related (Ran) regulates nucleocytoplasmic transport but also has a function in mitotic spindle formation and nuclear reformation after mitosis (Melchior, 2001). Being a GTPase, Ran hydrolyzes bound GTP into GDP, which alters Rans confirmation (Chook and Blobel, 2001). The GTP-bound Ran (RanGTP) is the active form, while the GDP-bound Ran (RanGDP) is the inactive form. Nucleocytoplasmic transport is dependent on the gradient of RanGTP/RanGDP, with RanGTP concentration being higher in the nucleus and RanGDP in the cytoplasm (Melchior, 2001). RanGTPase-activating protein 1 (RanGAP1) located on the cytoplasmic surface of NPCs and Regulator of chromosome condensation 1 (RCC1) (also called Ran guanine nucleotide exchange factor) located on chromatin in the nucleoplasm, maintain this gradient. In nucleocytoplasmic transport, Importin β (directly or via importin α) binds a nuclear localization sequence (NLS) on the transport cargo. Together, the cargo and Importin β are imported into the nucleus, through the central channel of the NPC. In the nucleus, RCC1 facilitates the release of GDP from Ran. Subsequently Ran binds GTP, as the cellular concentration of GTP is approximately ten times higher compared to GDP (Bos et al., 2007; Lui and Huang, 2009). RanGTP binds to Importin β with high affinity, which releases the cargo on the nucleoplasmic side (Jahed et al., 2016; Melchior, 2001). Importin β together with RanGTP is then exported from the nucleus to the cytoplasm. RanGAP1 anchored on the cytoplasmic side of the NPCs, facilitate the RanGTPase hydrolyzation of GTP to GDP (Melchior, 2001). This hydrolysis disrupts the binding of Importin β to Ran and Importin β is free to bind the next NLS containing cargo (Melchior, 2001). In the same way, cargos that are exported, from the nucleoplasm to the cytoplasm, contain a nuclear export signal (NES) that binds nuclear transport factors called exportins together with RanGTP. When this cargo:exportin:RanGTP complex reaches the cytoplasm, RanGAP1 facilitates the hydrolysis

15 of RanGTP, and the cargo with the NES, is released. RanGDP is transported back to the nucleus via binding to the nuclear transporter factor 2, which is released when RCC1 facilitates the release of GDP from Ran (Jahed et al., 2016; Lui and Huang, 2009).

The function of Ran in mitotic spindle formation is also linked to Importin β. Before mitosis, when a mother cell divides into two daughter cells, chromosomes are replicated into identical copies called sister chromatids. At the beginning of mitosis, the NE breaks down and the mitotic spindle is formed. When the NE breaks down, nuclear spindle-assembly factors (SAFs) forms inhibitory complexes with the cytoplasmic excess of Importin β (Barr and Gergely, 2007; Melchior, 2001). RCC1 is concentrated at chromatin and facilitates the release of GDP from Ran, which results in a high gradient of RanGTP at the chromatin surface. RanGTP binds Importin β which releases the SAFs that initiate nucleation of microtubules, resulting in the formation of the mitotic spindle on chromatin (Barr and Gergely, 2007; Lui and Huang, 2009). Many NE proteins, including the nuclear lamina and NPC proteins, are phosphorylated, which disrupts protein complexes and facilitates the clearance of the nuclear membrane from chromosomes (Liu and Pellman, 2020). At the centromeres on the chromosomes, assemble and connect to microtubules of the mitotic spindle. The two (the microtubule organization center also called spindle pole bodies) on opposite sides of the mitotic spindle, then pull the sister chromatids apart towards the opposite end of the cell and NE starts to reform. The cytoplasm then cleaves into two, with two genetically identical daughter cells formed. At the end of mitosis, the NE reassembles. Membrane elements locate around chromatin and fuse into a continuous nuclear membrane were Nups assemble into NPCs (Liu and Pellman, 2020). The high concentration of RanGTP around the mitotic chromosomes enables the dissociation of Nups from Importin β, which induces the formation of an NPC containing nuclear membrane (Walther et al., 2003). Even in the absence of chromatin, a continuous membrane with functional NPCs assembled around beads coated with RanGTP (Zhang and Clarke, 2001), whereas RanGDP-coated beads required RCC1 to form this membrane (Clarke and Zhang, 2008).

1.1.5 Chromatin In the nucleus, chromatin is organized and packed into large compact complexes containing both DNA and proteins (Kornberg, 1974). Euchromatin (loosely packed transcriptionally active chromatin) is mainly localized in the nuclear interior and underneath the NPCs (Bickmore and van Steensel, 2013; Fišerová et al., 2017), while heterochromatin (compact transcriptionally inactive chromatin) is enriched in the nuclear periphery but also at centromeres, telomeres and around nucleoli (Cremer and Cremer, 2010; Ritland Politz et al., 2016). DNA is wrapped around nucleosomes, octameric protein complexes consisting of two copies of the four core (H2A, H2B, H3, and H4). The linker histone 1 (H1) binds in between the nucleosomes which enable an even higher degree of compaction (Wolffe, 2001). Histones are small proteins with positively charged residues (arginines and lysines), especially at their N-terminal tails that can be post-translationally modified. Histones are among the most conserved proteins in (Alva et al., 2007; Butterworth, 2005). Some histones display different chromatin localization dependent on the histone variant. For example, there are four H3 variants in humans: H3.1, H3.2, H3.3, and CENP-A (Loyola and Almouzni, 2007; Szenker et al., 2011). H3.1 localize preferably to transcriptionally inactive regions, H3.2 show no preference, H3.3 localize preferably to sites with active or potentially active transcription (Ahmad and Henikoff, 2002; Delbarre et al., 2010; Stroud et al., 2012) and CENP-A localize to pericentromeric heterochromatin surrounding the centromeres (Black et al., 2004).

16

Histone variants and post-translational modifications Histones are incorporated into nucleosomes during DNA replication. However, H3.3 can also incorporate into nucleosomes in a replication-independent manner (Ahmad and Henikoff, 2002). The exact mechanism for this is not known, but three of the four substitutions between H3.1 and H3.3 (S87A, V89I, M90G), located in the central fold-domain, independently of each other, promotes incorporation in a replication-independent manner (Ahmad and Henikoff, 2002). The histone variants and the post-transcriptional modifications of histones, also known as the histone code, constitute an important part of the epigenetic transcription regulation (Jenuwein and Allis, 2001). Some modifications associate with transcriptionally active euchromatin, whereas others associate with transcriptionally inactive heterochromatin (Harr et al., 2016). Acetylation of the positively charged N-terminal tails of histones (for example acetylation of histone 3 at lysine 9 (H3K9Ac)) is associated with euchromatin (Kouzarides, 2007). The acetyl-group neutralizes the positively charged , which reduces the interaction with the negatively charged DNA backbone (Kouzarides, 2007, 2000). Enzymes called histone deacetylases (HDACs) remove acetyl-groups from histones (Seto and Yoshida, 2014) and histone acetyltransferases (HATs) add acetyl-groups on histones and thus alter the epigenetic state of chromatin (Kouzarides, 2000). Modifications such as methylation are more complex, with for example di- or tri-methylation of H3 at Lysine 4 (H3K4me2/me3) associated with euchromatin (Mohn et al., 2008) and methylation at Lysine 9 (H3K9me2/me3) associated with heterochromatin (Kouzarides, 2007).

Tethering chromatin to the nuclear periphery Gene-rich chromosome regions are separated from gene-poor regions in a tissue- specific way (Parada et al., 2004). For example, gene-rich chromosome 19 is located in the nuclear interior while gene-poor chromosome 18 is located in the nuclear periphery (Croft et al., 1999). Chromosome 5 localizes to the interior in liver cells but at the nuclear periphery in lung cells (Parada et al., 2004). This is thought to depend on the tissue-specific NE proteome, as overexpression of several tissue-specific NETs (Samp1, NET29, NET39, NET45, and NET47) independently repositioned parts of chromosome 5 to the nuclear periphery (Zuleger et al., 2013). This repositioning was dependent on the transmembrane domain of these NETs which indicate that these NETs, directly play a role in tethering chromosomes to the nuclear periphery. Out of these five previously mentioned NETs, only NET29 and NET39 promote a peripheral position of the gene-poor chromosome 13 (Zuleger et al., 2013), indicating that tissue-specific NETs tether chromatin in a tissue-specific way to the NE. Some of the INM proteins such as LBR and LEM-domain proteins (Lap2β, Emerin and Man1) directly bind chromatin (Berk et al., 2013; Nikolakaki et al., 2017), however most characterized INM proteins bind chromatin-binding proteins. Lap2β, Emerin, and Man1 bind BAF (Demmerle et al., 2012; Margalit et al., 2007; Zheng et al., 2000), and LBR binds HP1 (Amendola and van Steensel, 2014), which promotes the formation of heterochromatin (van Steensel and Belmont, 2017). Emerin and Lap2β also bind HDAC3 that removes acetyl-groups from H4, resulting in condensation of chromatin (Amendola and van Steensel, 2014; Berk et al., 2013; Zullo et al., 2012). The concentration of these heterochromatin-promoting proteins (BAF, HP1, and HDAC3) to the NE, results in gene repression at the nuclear periphery. Most genes are silenced when positioned at the INM (Francastel et al., 2001; Peric- Hupkes et al., 2010; Van de Vosse et al., 2011). Some NETs bind transcription factors, usually to sequester them, while others, such as SUN1, bind HATs which results in gene activation (Berk et al., 2013; Chi et al., 2007). Some genes are still active at the nuclear periphery (Cabal et al., 2006; Kim et al., 2011), which suggests that there are subdomains at the NE where genes

17 can be transcribed. For example, the NPCs are associated with transcriptionally active chromatin which could potentially facilitate the mRNA export of these genes and avoid chromatin condensation at the NPCs (which could obstruct nucleocytoplasmic transport) (Ibarra and Hetzer, 2015). Nups can also bind directly to promotor regions, some Nups even relocate to the nuclear interior to activate gene expression (Ibarra and Hetzer, 2015).

Chromatin organization during differentiation The NE is an incredibly complex structure that disassembles and reassembles, in every mitosis (Chen, 2012). The cell takes advantage of this NE disassembly-reassembly during differentiation, where a dramatic reorganization of the genome and NE proteome occurs (Kumaran and Spector, 2008; Zuleger et al., 2013). Chromatin gets more condensed and certain genes, not necessary for the cell's fate, become transcriptionally inactive (Gaspar-Maia et al., 2011; Ricci et al., 2015). Hi-C data, comparing human embryonic stem cells with their differentiated successors shows that around 36% of the genome switch between active and inactive compartments (Dixon et al., 2015). Multicellular life begins with two haploid gametes that fuse into a totipotent zygote that gives rise to a population of pluripotent cells, which differentiate into many different lineages. Pluripotent cells remain in their pluripotent stage due to core transcription factors such as Nanog, Oct3/4, and Sox2 which repress the expression of genes required for lineage commitment (Chambers and Tomlinson, 2009). When cells differentiate chromatin becomes more structured and condensed in comparison to their undifferentiated precursors, which have a more dynamic active transcriptome (Sivakumar et al., 2019; Talwar et al., 2013). Condensed chromatin regions can be divided into constitutive and facultative heterochromatin (Trojer and Reinberg, 2007). Constitutive heterochromatin contains highly repetitive DNA regions that are stably repressed, for example, pericentromeric heterochromatin and telomeres. While facultative heterochromatin is cell-type specific and only temporally inactive (Sivakumar et al., 2019). During differentiation, for example in myogenesis, neurogenesis, and adipocyte differentiation, there are several examples when tissue-specific genes, in facultative heterochromatin, move from nuclear periphery to the interior to be activated (Robson et al., 2016; Szczerbal et al., 2009; Williams et al., 2006). Chromatin interactions with lamins and the highly tissue-specific NE proteome that changes during differentiation could explain this tissue-specific spatial genome organization. The INM shows a large diversity with several hundreds of NETs expressed in a highly tissue- specific pattern (Korfali et al., 2012; Wilkie et al., 2011). In the last step of differentiation, when the cell goes into senescence (a state of irreversible growth arrest), constitutive heterochromatin repositions from the nuclear periphery to the interior into structures called senescence-associated heterochromatin foci (SAHF). These SAHFs prevent the proliferation of the cells and are a characteristic of aging (Parry and Narita, 2016).

18

1.2 Diseases of the nuclear envelope - Emery-Dreifuss Muscular Dystrophy

Proteins at the nuclear envelope have been associated with several human diseases, collectively called envelopathies or laminopathies (Burke and Stewart, 2014, 2002; Ho and Hegele, 2019). Most of the proteins found to cause these envelopathies are ubiquitously expressed, although the diseases have been associated with pathologies affecting specific tissues (Dauer and Worman, 2009). This is thought to depend on the highly tissue-specific NE proteome, were the tissue-specific NETs function together with the ubiquitously expressed NETs. These diseases are associated with both structural and functional defects of proteins at the nuclear envelope. The two main mechanisms thought to cause envelopathies are structural tension in tissues exposed to high mechanical stress, such as cardiac and skeletal muscles, and changes in gene expression due to abnormal chromatin organization (Bonne et al., 1993). Many of the envelopathies are directly or indirectly linked to mutations in LMNA or components of the nuclear lamina (Worman and Courvalin, 2005). Mutations in LMNA have been found in clinically distinct tissue-specific degenerative disorders called laminopathies. Laminopathies can be divided into four disorders, affecting striated muscles (heart and skeletal), adipose tissues, peripheral nerves, or several tissues such as premature aging syndromes (Progeria/HGPS) (see figure 3) (Dauer and Worman, 2009; Eriksson et al., 2003; Worman and Courvalin, 2005). Defects in the nuclear lamina destabilize the structure and rigidity of the whole nucleus making it more vulnerable to mechanical stress. The nuclear lamina also organizes chromatin at the nuclear periphery (see 1.1.1 The nuclear lamina), which is essential in the differentiation of many tissues including muscle, adipose and neural tissues (Batrakou et al., 2015; Maresca et al., 2012; Robson et al., 2016). The link between chromatin organization and disease is however not clear. For example, two different LMNA mutations (E161K, D596N) both causing cardiomyopathy, have the opposite effect on chromosome 13 localization. E161K repositions chromosome 13 to the nuclear interior while D596N tethers it closely to the nuclear periphery (Puckelwartz et al., 2011).

Figure 3, Laminopathies divided into four groups affecting different tissues (modified from (Dauer and Worman, 2009)).

19

The first envelopathy described, already in 1902, was Emery-Dreifuss Muscular Dystrophy (EDMD) (Cestan R, Lejonne NI, 1902). EDMD is a rare genetic muscle disorder, affecting around 1:100 000 people in the general population (Bonne et al., 1993; Norwood et al., 2009). EDMD is a muscular dystrophy clinically characterized by degeneration (atrophy) of certain muscles, fixed joints (contractures) and heart abnormalities (cardiomyopathy) (Madej-Pilarczyk, 2018; Morris, 2001). Initially, during late childhood, progressive muscle weakness and atrophy usually develops in the upper arms and lower legs although the onset, severity, and progression vary between EDMD patients. Generally, EDMD patients with early- onset have rapid disease progression and more severe complications than individuals with late- onset. Heart abnormalities often develop in early adulthood, with asymmetric heartbeats (arrhythmias), resulting in life-threatening progressive cardiomyopathy (Heller et al., 2020; Madej-Pilarczyk, 2018; Morris, 2001).

Most cases of EDMD are inherited as an X-linked or autosomal dominant (AD) disease. In X-linked EDMD, EDM (encoding Emerin) is mutated whereas most cases of the AD-EDMD are linked to mutations in LMNA (encoding for Lamin A/C) (Madej-Pilarczyk, 2018). AD- EDMD LMNA mutations have a broad variation in severity and onset, whereas X-linked EDMD, in general, have later onset and less acute cardiac problems (Madej-Pilarczyk, 2018; Morris, 2001). Emerin is ubiquitously expressed with high expression in skeletal and cardiac muscle and is depleted in approximately 95% of the X-linked EDMD patients (Berk et al., 2013; Yates et al., 1999). Emerin and Lamin A/C directly binds at the NE, with mutations in Lamin A/C often leading to the mislocalization of Emerin (Berk et al., 2013). Emerin and Lamin A/C are regulating muscle- and heart-specific gene expression and nuclear morphology, especially important in tissues exposed to mechanical stress such as skeletal and cardiac muscles (Guilluy et al., 2014; Holaska, 2008).

About half of the EDMD patients are connected to mutations in proteins (Emerin, Lamin A, 1/2, SUN1, and FHL1) known to cause EDMD when disrupted (Bonne et al., 1993; Heller et al., 2020; Madej-Pilarczyk, 2018). However, the disease mechanism is unclear as these proteins have functions in both mechanotransduction and gene regulation. Recently, 301 candidate genes were screened for mutations in 56 unlinked EDMD patients (Meinke et al., 2020). Many of the genes found mutated in these EDMD patients, are linked to gene regulation and the cytoskeleton, suggesting they are involved in the EDMD pathology (Meinke et al., 2020). The muscle-specific proteins found mutated (NET39, Tmem38A, Tmem214, WFS1, Samp1) also share the function of directly repositioning genes to the NE and potentially regulating gene expression during myogenesis (Meinke et al., 2020; Robson et al., 2016; Zuleger et al., 2013). As many of the proteins found mutated in these patients share common functions, this suggests that they work in common pathways important for muscle cells while causing EDMD when disrupted (Meinke et al., 2020).

20

1.3 The inner nuclear membrane protein Samp1

Samp1 was identified and characterized as an inner nuclear transmembrane protein by Buch and coworkers (Buch et al., 2009). Samp1 is also known as NET5 and Tmem201. Samp1 is homologous to NET5 in rats (Schirmer et al., 2003) and Ima1 in Schizosaccharomyces pombe (King et al., 2008) and CT.Samp1 in Chaetomium thermophilium (Vijayaraghavan et al., 2016). Samp1 has three isoforms, Samp1a (392 aa, 43 kDa) (Buch et al., 2009), Samp1b (566 aa, 62 kDa) (Borrego-Pinto et al., 2012) and Samp1c (666 aa, 75 kDa). The isoforms are produced by in exon 6, resulting in a non-spliced (SEKQP→FFPGD) unique sequence at the C-terminal (position 387-392) of Samp1a (see figure 4). In the N- terminal, a hydrophobic region (position 14-32) was originally predicted to be a transmembrane segment, although later proven to be nucleoplasmic (Gudise et al., 2011). Additionally, Samp1 has four conserved CXXC motifs in the nucleoplasmic N-terminal that potentially could form two zinc fingers. Samp1a and Samp1b has four transmembrane segments and Samp1c five transmembrane segments (see figure 4). The additional nucleoplasmic region between transmembrane 4 and 5, is very positively charged (pI around 10) and consists of 51 serine residues (19% of the region).

Figure 4: Membrane topology of Samp1 isoforms. Samp1a comprises aa 1-392 ending with a FFPGD pentapeptide (gray). Samp1b comprises of aa 1-566 and Samp1c comprises of aa 1-666, both with SEKQP instead of the unique pentapeptide of Samp1a. The nucleoplasmic N-terminal domain contains a hydrophobic sequence (blue) and four conserved CXXC motifs (yellow) (Paper V).

21

Samp1 in Emery-Dreifuss Muscular Dystrophy Samp1 was found to be mislocalized in AD-EDMD patients (Mattioli et al., 2018) and to detach the centrosome from the NE when depleted (Gudise et al., 2011), a common phenotype in EDMD patient cells (Salpingidou et al., 2007). This year, substitution mutations were found in Samp1 (G15A, G18S, G595S) in combination with other proteins, in three EDMD patients (Meinke et al., 2020). Two of these mutations (G15A, G18S) is in the hydrophobic part of the N-terminal. The other mutation (G597S) is at the positively charged part between transmembrane 4 and 5 in Samp1c. One of the patients had an unusual mutation in LMNA, not previously associated with muscular dystrophies, suggesting both mutations (Samp1 G15A, LMNA R249Q) contribute to a more severe EDMD phenotype with early-onset (1 year). In the two other patients, mutations in LINC complex, cytoskeleton and sarcolemma proteins were combined with the Samp1 mutations (G18S, G595S) giving rise to different onset and severity (Meinke et al., 2020).

Samp1 and RanGTPase Samp1 coprecipitates with Ran both in and mitosis (Jafferali et al., 2014). This interaction was later shown to be direct, with a preference for Samp1 to bind RanGTP over RanGDP (Vijayaraghavan et al., 2016). In two genome-wide siRNA screens in (Sönnichsen et al., 2005) and HeLa cells (Neumann et al., 2010) Samp1 depletion was shown to cause mild mitotic effects. Samp1 is the first membrane protein found located in the mitotic spindle hence the name Spindle-Associated Membrane Protein 1 (Buch et al., 2009). Later two other NETs; Tmem2014 and WFS1 were also shown to localize to the mitotic spindle (Wilkie et al., 2011). Samp1 has a function in stabilizing the mitotic spindle (Larsson et al., 2018). The mitotic spindle is formed by nucleation of microtubules from centrosomes and chromosomes (Goshima et al., 2008; Meunier and Vernos, 2012). The Augmin complex recruits γ-tubulin to pre-existing microtubules, where these proteins initiate nucleation and branching of microtubules (Goshima et al., 2008). Samp1 directly binds to γ- tubulin and interacts with Haus6 of the Augmin complex and is thought to recruit these SAFs to the mitotic spindle. Samp1 depletion prolonged metaphase and caused a less compact and less organized metaphase plate, resulting in chromosome mis-segregation, which indicates that Samp1 together with the Augmin complex and γ-tubulin has a stabilizing role on the mitotic spindle (Larsson et al., 2018). Unlike many other G-proteins, Ran lacks both lipid anchors and amphipathic α-helixes necessary for interactions with membranes. As Samp1 is Rans only known transmembrane binding partner, Samp1 is thought to provide a binding site at the INM for RanGTP to facilitate the execution of its many functions locally (Vijayaraghavan et al., 2016). Also in in the mitotic spindle, Samp1 might have the role of providing a local binding site for RanGTP at the site of nucleation on pre-existing microtubules (Larsson et al., 2018).

Tissue-specific expression of Samp1 Samp1 is highly expressed in neuronal, muscle and testicular tissue (Figueroa et al., 2010; Thanisch et al., 2017; Worman and Schirmer, 2015). Membrane cross-linking immunoprecipitation (MCLIP) of U2OS cells shows that Samp1 interacts with SUN1, Emerin, and Lamin B1 (Jafferali et al., 2014). Samp1 has also been shown to interact with SUN2 and Lamin A/C in mouse NIH3T3 fibroblasts (Borrego-Pinto et al., 2012). In muscle cells, Samp1 has an essential role in myogenesis (Jafferali et al., 2017). Furthermore, muscles are highly exposed to mechanical stress and require a functional LINC complex. During neuronal differentiation, nuclear migration is an essential part. Samp1 was shown to be a part of the TAN lines and involved in nuclear migration (Borrego-Pinto et al., 2012) and it is tempting to speculate that Samp1 has a role in this differentiation also.

22

2. Aim

There are several hundreds of different nuclear envelope transmembrane proteins, many of which display a tissue-specific expression pattern. Most of them are uncharacterized, but a few of them have been linked to a diverse group of human disorders collectively called envelopathies or laminopathies. In this thesis, the overall aim was to investigate the functional organization of the NE concerning chromatin organization and cell differentiation.

The specific aims of this thesis were to:

• Develop an image analysis tool to enable visualization of epigenetic spatiotemporal

changes in chromatin organization at the nuclear periphery in live cells.

• Investigate the roles of Samp1 in chromatin organization and cell differentiation.

• Characterize the functional interaction of Samp1 with Ran and Emerin.

• Investigate Emery-Dreifuss Muscular Dystrophy mutants of Samp1.

23

3. Methodological considerations

3.1 Cell types In paper I, II, and V, we used U2OS cells derived from bone tissue of a fifteen-year-old girl suffering from osteosarcoma in 1964 (Ponten and Saksela, 1967). U2OS cells have an epithelial-like morphology and contain a large nucleus suitable for image analysis. These cells are contact inhibited and form a uniform nonproliferating population when reaching confluency. This results in a more similarly structured chromatin distribution were smaller changes could be detected (paper I). Chromosome counts are in the hypertriploid range according to ATCC (U-2 OS ATCC HTB-96). However, all 10 monoclonal KO colonies picked, using the five different sgRNAs (see 3.5 CrisprCas9 genome editing) were completely depleted of Samp1. In paper II, we used tsBN2 cells derived from BHK21 (Baby hamster kidney cells) carrying a temperature-sensitive mutation in RCC1, which inactivates RCC1 at the restrictive temperature 39.5 °C (Nishitani et al., 1991). This results in a failure to regenerate RanGTP, as RCC1 is the only guanine nucleotide exchange factor that can release GDP from Ran (see 1.1.4 RanGTPase in nucleocytoplasmic transport and mitosis) (Jahed et al., 2016). In paper III, we used human induced pluripotent stem cells (iPSCs) ATCC ACS-1011 derived from human foreskin fibroblasts as a model for differentiation. In general, iPSCs express the same markers as embryonic stem cells (ESCs) and show similar morphology (Chin et al., 2009; Constantinescu et al., 2006; Takahashi and Yamanaka, 2006). We noticed that these cells were sensitive to vibrations and matrigel concentration, which were optimized to prevent spontaneous differentiation. Subculturing was done weekly using a micromanipulator to scratch a section from the monolayer center of the iPSC-colony to isolate undifferentiated cells from potential differentiated peripheral cells. In paper IV, PC6.3 and SH-SY5Y cells were used as a model for neuronal differentiation. PC6.3 cells (a subline of PC12 cells) are chromaffin cells of the rat adrenal medulla (Mills et al., 1995). In response to nerve growth factor (NGF) these cells extend neurites and stop dividing. These neuronal-like cells become dependent on NGF to survive (Pittman et al., 1993). SH-SY5Y cells (a subline of SK-N-SH) are neuroblastoma cells from a neuroendocrine tumor of a four-year-old girl. In response to retinoic acid (RA), these cells differentiate into a cholinergic-like phenotype with extended neurites (Forster et al., 2016). In paper IV and V, we used HeLa cells derived from Henrietta Lacks in 1951 who suffered from cervical cancer. HeLa cells are the most widely used human cancer cell line today (Masters, 2002). These cells contain a large nucleus suitable for image analysis using FRIC.

3.2 Live cell imaging In paper I and IV, we use live-cell imaging. In conventional immunofluorescence (IF) microscopy of fixed samples, images can only be captured at one certain time-point and the fixation procedure may alter cellular structures. There are also limitations concerning resolution, antibody specificity and accessibility. Live cell imaging allows measuring dynamic changes during chosen time-periods. To avoid some of the limitations with conventional IF microscopy, we developed FRIC for live-cell imaging by using fluorescent protein (FP) -tagged histones (see 3.3 FRIC below).

24

3.3 Fluorescent Ratiometric Imaging of Chromatin (FRIC) FRIC uses Histone 3.3-EGFP as a marker for euchromatin and H2B-mCherry as a marker for general chromatin. These FP-tagged histones have previously been shown to behave as endogenous histones (Delbarre et al., 2010; Kimura and Cook, 2001). FP-tagged H3.3 correlates with markers for transcriptionally active chromatin (Ahmad and Henikoff, 2002; Delbarre et al., 2010; Lin et al., 2013). Using chromatin immunoprecipitation (ChIP), FP- tagged H3.3 correlated with active chromatin and transcribed or potentially transcribed regions (Delbarre et al., 2010; Tamura et al., 2009). FP-tagged H3.3 is post-translationally modified as wt H3.3 (Delbarre et al., 2010). FP-tagged H2B shows no preference for active or repressed chromatin and is thus a good marker for general chromatin. Minor fractions of FP-tagged H3.3 have also been found in repressed genomic regions (Delbarre et al., 2010), for example, constitutive heterochromatin regions associated with telomeres and centromeres which are stably repressed (Ivanauskiene et al., 2014; Udugama et al., 2015). In paper I, we developed the FRIC method to monitor chromatin organization in live cells. A vector (pTandemH) with two separate multicloning sites with an internal ribosome entry site in between, allowed insertion of both H2B-mCherry and H3.3-EGFP into the same vector with expression controlled by the same promotor. This enables stoichiometric constant expression of the two FP-tagged histones. To normalize the intensities, intra- and intercellularly, we divided the channels by their mean value and variance. This allows us to divide the two channels to generate a ratio image (normalized H3.3-EGFP/normalized H2B- mCherry). The cells were equilibrated for at least 48h post-transfection, as H3.3 and H2B have different incorporation kinetics in chromatin. H3.3 incorporates both during replication and in a replication-independent manner while H2B is only incorporated during replication. The ectopic expression of both these FP-tagged histones has previously been studied without signs of toxicity and abnormalities (Delbarre et al., 2010; Kanda et al., 1998). To ensure that the intensities were at appropriate levels an ImageQuality module was added to the CellProfiler pipeline. If 0.2% of the pixels in the nucleus were saturated/underexposed or the Otsu-threshold value was above 0.15, the images were removed. The Otsu-threshold divides the image into two classes, trying to minimize the variance between the foreground and background. High Otsu-threshold results in bad segmentation of the nucleus possibly due to the disturbing background. All nuclei were also checked manually for abnormal morphology and in such cases removed from the set. The ratiometric images were divided into 40 concentric zones with equal width and consequently, the peripheral zones contain more pixels and have higher significance in comparison to the inner nuclear zones. Dividing the nuclear area into four zones with equal area gave similar results but with a lower resolution in the nuclear periphery (see paper I). Mathematically, the pixels of the diameter of the nucleus was not large enough to divide into more than four zones with equal area, as it is desired that the outer zone should include a width of at least two pixels. Z-stacks with a slice thickness of 0.5 µm were collected. The confocal optical z-section containing most pixels were considered equatorial and used for further analysis with FRIC. The grazing sections were compared to the equatorial section and showed a similar trend with more heterochromatin in the periphery, but the profile was less pronounced with less difference to the interior zones (see paper I).

3.4 Fluorescence recovery after photobleaching (FRAP) In paper II, we performed fluorescence recovery after photobleaching (FRAP) experiments to determine the mobility of FP-tagged transmembrane proteins in the NE. In FRAP, an area of the cell is photobleached, for example, a part of the nuclear envelope, then

25 the halftime (t1/2) of the fluorescence recovery was measured. In paper II, t1/2 represents the time it takes for half of the photobleached YFP-Emerin to be exchanged with non-bleached YFP-Emerin and hence the mobility of YFP-Emerin. As most cells migrated during the experiment, the ImageJ plugin StackReg was used to segment out the nucleus, and each time- point was normalized to the time-points before and after, to get a corrected intensity value of the chosen FRAP area.

3.5 Gene knock out using CRISPR/Cas9 mediated genome editing In paper I, II, and V, we used CRISPR/Cas (clustered regularly interspaced short palindromic repeats/CRISPR associated system) genome editing based on an RNA-guided nuclease Cas9. Cas9 recognizes double-stranded DNA (dsDNA) at the PAM sequence 5´- NGG-3´ and thereafter binds the complementary sgRNA (upstream of the PAM) (Sampson and Weiss, 2014). The two endonuclease domains (RuvC-I and HNH) of Cas9, then cleave both strands of the target sequence four nucleotides upstream of the PAM sequence. The cell repairs this break by homologous direct repair (HDR) or nonhomologous end-joining (NHEJ). HDR only occurs in the presence of homologous sequences that could be used as a template for the repair. In NHEJ, one nucleotide pair is cleaved off and a frameshift in the target gene introduced, which usually leads to a premature stop codon (Sampson and Weiss, 2014). Five different sgRNAs were designed to target the first exon of the human Samp1 gene. sgRNA 1-4 was designed using the E-CRISPR online application (German Cancer Research Center (DKFZ), 2016) and sgRNA5 was designed with the Zhang Lab online CRISPR design and analysis application (Zhang Lab, 2016). The sgRNAs with the highest score (meaning lowest off-target probability) were chosen and a scrambled sequence with low homology to the human genome was chosen as a control. sgRNA4 was designed to work in both the human and rat Samp1 gene.

3.6 Microscale thermophoresis (MST) In paper II, we used microscale thermophoresis (MST) to provide close-to-native conditions for protein interactions in solution. MST uses an infrared laser to induce a temperature gradient and measures the movement of bound and unbound proteins in a dose- response curve, from where KD values can be derived (Jerabek-Willemsen et al., 2011). The concentration of fluorescently labeled recombinantly expressed CT.Samp1(1–180) comprising the nucleoplasmic part of Samp1 from Chaetomium thermophilum, was kept constant while GST-Emerin and GST-Ran concentrations were varied. Special MST premium capillaries were used to avoid non-specific binding to the capillary wall and GST was used as a negative control. The Samp1 homolog of the thermophilic fungus Chaetomium thermophilum (CT.Samp1) was used as the solubility of the recombinantly expressed protein was better compared to the human counterpart.

3.7 Super-resolution using Structured Illumination Microscopy (SIM) In paper I, we used structured illumination microscopy (SIM) to achieve a higher resolution. We used 3D SIM with three excitation beams that created a pattern using grids, with both lateral and axial components. The point spread function was then derived before the images were processed. The resolution was calibrated and measured using 40 nm green fluorescent beads, to the maximum precision of 102 ±5 nm laterally and 250 ±12 nm axially, which is better compared to confocal microscopy with an approximately maximum resolution of 200 nm laterally and 500 nm axially.

26

4. Results and Discussion

Monitoring chromatin organization at the nuclear periphery (Paper I) In the first part of paper I, we developed an image analysis tool called Fluorescence Ratiometric Imaging of Chromatin (FRIC). FRIC is used to quantitatively monitor the dynamic spatiotemporal distribution of euchromatin and total chromatin in live cells. A tandem vector (pTandemH) enables stoichiometric expression levels of the two histone variants, Histone 3.3 fused to EGFP and Histone 2B fused to mCherry. H3.3-EGFP serves as a marker for transcriptionally active chromatin (euchromatin) while H2B-mCherry serves as a marker for general chromatin (Delbarre et al., 2010). Ratiometric images of H3.3-EGFP/H2B-mCherry correlated better than H3.3 alone, with the euchromatin marker RNA polymerase II and less with the heterochromatin marker H3K9me3, meaning that the ratio served as a better marker for euchromatin than simply using H3.3. The ratiometric images were divided into concentric zones of equal width and the mean values plotted as radial profiles from the nuclear periphery to the nuclear interior. The radial profile displayed a concentrated distribution of heterochromatin in the nuclear periphery of U2OS cells, consistent with studies using other techniques (Davies, 1967; Fišerová et al., 2017; Gruenbaum and Foisner, 2015; Lieberman-Aiden et al., 2009; Peric-Hupkes et al., 2010). As proof of concept that FRIC accurately measures the epigenetic state of chromatin, peripheral heterochromatin responded to experimental manipulation using agents that affect histone acetylation. As expected, cells treated with the HAT (histone acetyltransferase) inhibitor Anacardic acid (AA), which then decreases acetylation of histones, resulted in more peripheral heterochromatin (lower ratio). In contrast, the HDAC (histone deacetylase) inhibitor Trichostatin A (TSA), which then decreases the deacetylation of histones, resulted in an opposite decrease of heterochromatin (higher ratio). As a control, we used the heterochromatin marker H3K9me3 to verify that the HDAC inhibitor resulted in less heterochromatin in the nuclear periphery compared to control cells. Expression of progerin, Lamin A L647R, a mutation that disrupts the ZMPSTE24 cleavage site in prelamin A, causing the disease Hutchinson-Gilford Progeria Syndrome (HGPS) (Wang et al., 2016), was used as a disease model (see 1.1.1 Lamin A processing). Cells expressing progerin had substantially less heterochromatin in the nuclear periphery. Together, the results strongly suggest that FRIC correctly monitors dynamic chromatin organization events at the nuclear periphery. After treating cells with the HDAC inhibitor TSA, the outermost peripheral zone in the treated cells revealed a significant decrease in heterochromatin already after one hour, illustrating the usefulness of FRIC in monitoring dynamic reorganization of chromatin in live cells. To be able to visualize changes in chromatin organization at an even more detailed level, we showed that FRIC was compatible with structured illumination microscopy (SIM) giving higher image resolution. Already after analyzing a few cells, we noticed a difference in the first 10 zones in comparison to 5 zones using the conventional LSM780 confocal microscopy. We also showed that confluent U2OS cells (which stop proliferating) formed a more uniform population with more organized chromatin compared to proliferating cells (that in contrast represent many different stages in the cell cycle). Although synchronization of cells may enable a higher sensitivity, we showed that the results could be replicated in proliferating cells, suggesting that FRIC can be applied to monitor changes in chromatin organization also in asynchronous cultures. In conclusion, FRIC is a versatile and straight forward, easy to use, tool that enables visualization and quantification of spatial epigenetic chromatin reorganization events in single live cells. FRIC is an unbiased method in comparison to crosslinkers or immunostaining where

27 antibodies suffer from accessibility in compact heterochromatin structures. FRIC is not designed to register global appearance/disappearance of euchromatin/heterochromatin, but rather to measures the relative distribution of the epigenetic state of chromatin, especially in the nuclear periphery.

The effect of Samp1 on chromatin distribution (Paper I, IV and V) Samp1 overexpression has been shown to alter chromosome localization by repositioning parts of chromosome 5 to the nuclear periphery in human fibrosarcoma cells (Zuleger et al., 2013). In yeast, Ima1 (Samp1 homolog) is associated with the centromeres but not with telomeres or the pericentromeric heterochromatin surrounding the centromere (Steglich et al., 2012). The depletion of Ima1 leads to inefficient tethering of the centromeres to the nuclear periphery (King et al., 2008). Thus, we investigated if Samp1 influenced chromatin organization using FRIC. We depleted Samp1 using knockdown (KD) by RNAi and knock out (KO) by CRISPR/Cas9 targeted genome editing. In paper I, U2OS cells depleted of Samp1 showed a significantly higher ratio (H3.3/H2B) in the nuclear periphery compared to controls, which indicates less heterochromatin at the nuclear periphery. Overexpression of Samp1 resulted in an opposite lower ratio at the nuclear periphery, indicating more heterochromatin. In paper IV, PC6.3 cells depleted of Samp1 (KO-Samp1) also had a significantly higher ratio in the nuclear periphery compared to the control PC6.3 cells. Taken together the results show that Samp1 promotes peripheral heterochromatin. As previously mentioned (see 1.1.5 Chromatin), other INM proteins such as the LBR and LEM-proteins directly bind chromatin or heterochromatin-promoting proteins which results in a repression of chromatin at the nuclear periphery. The effect of Samp1 on peripheral heterochromatin could consequently be direct or indirect through its binding partners in the NE. For example, the depletion of Emerin displays more euchromatin in the nuclear periphery similar to that of HDAC3 depleted cells (Bhaskara et al., 2010). In paper V, we show that the soluble nucleoplasmic N-terminal domain of truncated Samp1 (1-180), without its transmembrane domains, colocalizes with heterochromatin marker H3K9me3 without its binding partners in the NE, suggesting that Samp1 may directly interacts with chromatin.

Interactions between nuclear envelope proteins (Paper II) Samp1 binds directly to Emerin (in a zinc-dependent manner) (Jafferali et al., 2014) and the Ran-GTPase (Vijayaraghavan et al., 2016) and hypothetically provides an anchoring site to maintain the high RanGTP-gradient close to the nuclear periphery (see 1.1.4 RanGTPase in nucleocytoplasmic transport and mitosis). However, several other binding partners in live cells could affect the binding between Samp1 with Emerin and Ran. Thus, in paper II, we further investigated the direct binding of Samp1 to Ran and Emerin. To do this we generated monoclonal U2OS cells depleted of Samp1 (see 3.5 Gene knock out using CRISPR/Cas9 mediated genome editing). FRAP experiments showed that the KO-Samp1 cells displayed increased mobility of YFP-Emerin in contrast to cells overexpressing Samp1 which displayed lower mobility of YFP-Emerin, in the NE. The results show that Samp1 significantly reduces the mobility of Emerin in the NE. tsBN2 cells carrying a temperature-sensitive mutant of RCC1 (the guanine nucleotide exchange factor of Ran) were cultured at the restrictive temperature (39.5 °C) to deplete RanGTP (see 3.1 Cell types). In the RanGTP depleted cells, Emerin had lower mobility suggesting that the mobility of Emerin is dependent on RanGTP. Using MST, we observed that Samp1 had a high affinity for both Emerin and Ran. This high-affinity binding between Samp1- Emerin is comparable with Emerins high-affinity binding to Lamin A and BAF, showing that Samp1 also is an important interaction partner of Emerin. In the presence of RanGTP, the 28 affinity between Samp1 and Emerin decreased suggesting that RanGTP increases the mobility of YFP-Emerin by reducing the interaction between Samp1 and Emerin. In paper II, we further show that the binding site of Emerin is located between the amino acids 1-79 of Samp1. The RanGTPase binds a domain containing the amino acids 80-180 of Samp1, which preferentially binds Ran-GTP over Ran-GDP (Vijayaraghavan et al., 2016). The different binding domains of Ran and Emerin on Samp1 and the lack of binding between Ran and Emerin, suggest that this regulation is allosteric. It is tempting to speculate that RanGTP interrupts the Samp1-Emerin binding in a similar way in which RanGTP interrupts the binding between the import cargo and Importin β (see 1.1.4 RanGTPase in nucleocytoplasmic transport and mitosis). Emerin is a relatively dynamic NET and its mobility at the NE was substantially affected by the Samp1 expression level (paper II). Samp1 expression is tissue-specific and especially high in muscles, whereas Emerin is ubiquitously expressed. Our results provide a good illustration of how the properties of widely expressed NETs may differ in different tissues. This may give a clue to the seeming paradox that tissue-specific laminopathies are linked to widely expressed NE proteins (Worman and Schirmer, 2015).

Samp1 in cell differentiation (Paper III and IV) Samp1 is essential for myoblast differentiation (myogenesis) in mouse C2C12 cells, whose Samp1 expression level increased 7-fold during differentiation (Jafferali et al., 2017). In paper III, we used human induced pluripotent stem cells (iPSCs) as a model to study stem cell differentiation. These cells were either cultured under pluripotent conditions or in differentiation medium and classified through established morphological characteristics and immunostaining of the early differentiation marker Lamin A/C. During iPSC differentiation, the Samp1 levels increased at the nuclear periphery (simultaneously with Lamin A/C expression), while Emerin and Lamin B1 levels persisted. The increased Samp1 levels in the nuclear envelope in differentiated iPSCs raised the question of whether Samp1 could affect the differentiation process. During pluripotent culturing conditions, ectopic expression of YFP-tagged Samp1 induced fast differentiation of the entire iPSC colony, whereas YFP-Emerin or YFP did not, indicating that Samp1 promotes differentiation. Several of the cells in the Samp1-induced differentiated colonies were positive for βIII- tubulin (a marker for neuronal differentiation) already after 6 days post-transfection despite pluripotent culturing conditions. This was earlier in comparison to cells cultured in differentiation medium that had comparable βIII-tubulin levels first after 15 days. Thus, we conclude that Samp1 induced a remarkably fast differentiation of iPSCs. Samp1 is highly expressed in neurons and associated with the LINC complex that is essential for nuclear migration, which is important for the development of the central nervous system (see 1.1.2 The function of the LINC complex). The increase in βIII-tubulin levels during the Samp1-induced differentiation of the hiPSCs raised the question if Samp1 had a function in neuronal differentiation. In paper IV, we made efforts to answer this question however the results showed that depletion of Samp1 did not affect the differentiation of neuronal PC6.3 cells. These KO-Samp1 cells gained extended neurites similar to the control cells, after 5 days of NGF treatment. PC6.3 cells are however at a late stage of neuronal differentiation compared to the pluripotent stem cells. It would be intriguing to see if the depletion of Samp1 prevents cell lineage (such as neuronal and muscle) differentiation of iPSCs.

29

Monitoring chromatin organization during differentiation (paper IV) In paper IV, we present an alternative way to display FRIC. By using the 75th percentiles of the ratio (normalized H3.3-EGFP/H2B-mCherry) and the inverse ratio (normalized H2B- mCherry/ H3.3-EGFP) followed by merging the two images, it becomes easier to see where heterochromatin and euchromatin is distributed in the nucleus. Treating HeLa cells with the HAT inhibitor TSA or the HDAC inhibitor AA resulted in decreased and increased distribution of the 75th percentile of the inverse ratio in the nuclear periphery, respectively. A similar effect on the epigenetic state of chromatin was also shown for U2OS cells in paper I, although this was displayed using the radial profile and relative ratio (nuclear periphery/nuclear interior). We further showed in paper IV, that the FRIC representation using the percentiles could be used to monitor chromatin reorganization during differentiation. SH-SY5Y and PC6.3 cells were differentiated for 5 days and compared to undifferentiated cells. Differentiated cells displayed a lower ratio in the periphery meaning more peripheral heterochromatin in comparison to the undifferentiated cells. SH-SY5Y and PC6.3 cells are neuronal cells at a late stage of differentiation in contrast to most other studies on ESCs and iPSCs that previously have been used to show that chromatin becomes redistributed and more condensed during differentiation. It is therefore remarkable that we could see this subtle epigenetic change in chromatin organization at this relatively late stage of differentiation. Samp1 depleted (KO-Samp1) PC6.3 cells had a higher peripheral ratio compared to control cells, showing that Samp1 also promotes peripheral heterochromatin in PC6.3 cells as in the U2OS cells (paper I). However, no significant increase in peripheral heterochromatin was seen during the differentiation of these KO-Samp1 PC6.3 cells, even though these cells differentiated as the control cells. However, there are several hundreds of NETs expressed in a tissue-specific manner which together with Samp1 could function in neuronal differentiation. It remains to be investigated whether Samp1 might play a more important role in earlier stages of neuronal differentiation.

Emery-Dreifuss Muscular Dystrophy mutations of Samp1 (paper V) Three novel substitution mutations of Samp1 were found in patients suffering from Emery-Dreifuss Muscular Dystrophy (EDMD) (Meinke et al., 2020). In paper V, we investigated two of the mutations (p.G15A and p.G18S) located in the hydrophobic part of the nucleoplasmic N-terminal tail, which is present in all Samp1 isoforms (see 1.3 The inner nuclear membrane protein Samp1). The mutants YFP-Samp1a p.G15A and YFP-Samp1a p.G18S localized at the NE as YFP-Samp1a, where no effect on nuclear morphology or size could be detected. Samp1 is a part of the transmembrane actin-associated nuclear (TAN) lines, which are essential for positioning the centrosome in front of the nucleus during nuclear migration (Borrego-Pinto et al., 2012). As previously described (Buch et al., 2009), depletion of Samp1 results in centrosome detachment from the NE, a common phenotype in EDMD patient cells (Salpingidou et al., 2007). Thus, we investigated if the EDMD mutants of Samp1 could influence the distance between the centrosome and nucleus after overexpression of the two mutants. Although overexpression of wt YFP-Samp1a resulted in a larger distance between the centrosome and the NE compared to control cells, YFP-Samp1a p.G15A or YFP-Samp1a p.G18S had no additional effect. The depletion of Emerin from the NE is one of the most common causes of EDMD (Heller et al., 2020; Yates et al., 1999). The mobility of Emerin in the NE depends on the direct binding to the first 1-79 amino acids of Samp1 (Jafferali et al., 2014; paper II). Thus, we investigated if the mutations in the N-terminal affected Emerin localization, as a possible lead to explain these mutations in EDMD. However, in comparison to the overexpression of wt YFP-Samp1a, there was no additional effect of the mutants on Emerin localization.

30

We previously showed that Samp1 promotes peripheral heterochromatin (paper I and paper IV), although it was unclear whether this effect was indirectly mediated via its interaction partners in the NE or via a direct interaction with chromatin. Therefore, we designed and expressed the nucleoplasmically exposed part of Samp1 (without its transmembrane segments). YFP-Samp1 (1-180) distributed in the nucleoplasm and accumulated in perinucleolar regions and colocalized with the heterochromatin marker H3K9me3 and compact DNA, visualized by Draq5. The results suggest that Samp1 interacts with heterochromatin.

31

5. Conclusions

• A novel method called Fluorescent Ratiometric Imaging of Chromatin (FRIC) was developed, which enabled visualization of epigenetic spatiotemporal changes in chromatin organization at the nuclear periphery of live cells. • Samp1 promotes peripheral heterochromatin and might interact with chromatin independent of its neighboring binding partners at the NE. • The mobility of Emerin in the NE depends on the presence of its binding partner Samp1 and this interaction was regulated by RanGTP. • Overexpression of Samp1 induced rapid differentiation of the entire iPSC colonies with a fast expression of the neuronal marker βIII-tubulin. • FRIC detects reorganization of chromatin at the nuclear periphery during neuronal differentiation. • EDMD mutations in Samp1 had no significant effect on targeting to the INM, Emerin distribution, centrosome localization, nuclear size or morphology.

32

6. Future Investigations

FRIC is an easy-to-use, versatile and unbiased method that requires a minimum of experimental steps and therefore is suitable for screening purposes. A monoclonal cell line expressing pTandemH would be a great asset for screening, as this limits the variance in expression levels and increases the numbers of nuclei captured with fluorescent signals that could be used in the further imaging analysis. FRIC has a great potential for screening and could be used with confocal, wide-field or super-resolution microscopy. Effects on chromatin organization using a known histone deacetylation agent were detected already after 1h, which potentially could open-up for screening for other potential chromatin-modifying agents or proteins. Genome-wide screening of NET specific RNAi libraries could easily be implemented for high-throughput format using FRIC.

An interesting experiment would be to deplete Samp1 or other tissue-specific NETs in hiPSCs, to investigate if these NETs are essential in different tissue-specific differentiation. Although hiPSCs are hard to transfect, lentiviral transduction could be used. As FRIC enables us to monitor chromatin organization in live cells it would be interesting to see at what stage of differentiation, the epigenetic state of chromatin changes. For example, as Samp1 is essential for myogenesis, it would be interesting to see if the epigenetic state of chromatin changes even without Samp1 expression when the myoblasts are treated with differentiation medium.

Samp1 is essential for the maintenance of peripheral heterochromatin, however, this effect could be indirect through its binding partners at the nuclear periphery. The truncated nucleoplasmic part of Samp1, without its transmembrane domains, colocalizes with heterochromatin, which suggests that Samp1 binds chromatin directly. Co- immunoprecipitation of Samp1 with histones and heterochromatin markers or DNA-binding assays could clarify if Samp1 interacts directly with DNA, chromatin-binding proteins, or both.

So far overexpression of the EDMD Samp1 mutants has not revealed any obvious defects or phenotypic alterations that could be related to the EDMD disease mechanism. On a previous study overexpression of shRNA resistant wt Samp1a was able to rescue muscle cell differentiation in Samp1 depleted cells (Jafferali et al., 2017). Using a similar approach, it would be interesting to determine if and to what extent these mutants can rescue myotube formation. Furthermore, the effects of wt and mutant Samp1 on chromatin organization can be determined using FRIC as readout and their potential chromatin binding properties could be analyzed.

33

34

7. Populärvetenskaplig sammanfattning

En , så som människan, består av olika sorters celler som tillsammans bygger upp vävnader och organ i vår kropp. Livet börjar med att två könsceller slås ihop till en cell. Denna cell kopierar sig själv och differentierar till alla specialiserade vävnader och organ i vår kropp. När denna differentiering sker, så blir vissa delar av det genetiska materialet (arvsmassan) inaktivt. Alla celler har exakt samma genetiska material men det uttrycks olika. Då en muskelcell inte har samma funktion som en nervcell, kan de nervspecifika generna vara inaktiva i en muskelcell. Denna reglering kontrolleras av olika proteiner i cellerna. Våra celler har en cellkärna, där det genetiska materialet DNA finns. Cellkärnan har två membranväggar, det yttre och det inre membranet. Under dessa membran finns cellkärnans skelett som ger kärnan dess struktur. Detta skelett reglerar också det genetiska materialet genom att främst inaktivera gener vid cellkärnans membran. I det inre membranet finns det flera hundra olika proteiner. Beroende på vilken vävnad eller funktion en cell har, behövs det olika kärnmembran- proteiner. Några av dessa finns i alla celler medan andra bara finns i några specifika celler så som muskelceller.

I denna doktorsavhandling så utvecklade vi en teknik som vi kallar fluorescens ratiometrisk avbildning av kromatin (FRIC). I FRIC så uttrycker vi två färgade proteiner som binder det genetiska materialet. Det ena proteinet (Histon 3.3) binder till aktiva regioner på det genetiska materialet medan det andra proteinet (Histon 2B) binder både aktiva och inaktiva regioner. Vi kan genom att ta bild på dessa två proteiner i mikroskop se vart det finns aktiva regioner i cellkärnan. Vi använde oss av FRIC för att se vad som skedde när vi tillsatte olika kemikalier till cellerna. Vi fick mer eller mindre inaktiva delar av det genetiska materialet vid det inre kärnmembranet. När vi differentierade nervceller till att bli ännu mer specialiserade, så kunde vi se att mer av det genetiska materialet blev inaktiverat vid membranet. Vi studerade ett av proteinerna i det inre membranet med namnet Samp1. Vi har i denna avhandling visat att Samp1 hjälper cellkärnans skelett att bilda inaktiva regioner av det genetiska materialet vid det inre kärnmembranet. Vi visade också att Samp1 binder till ett annat protein vid namn Emerin i det inre membranet och att denna bindning regleras av den aktiva formen av proteinet Ran. Människor som saknar proteinet Emerin i sina celler, får sjukdom Emery–Dreifuss muskeldystrofi. Förändringar (mutationer) i Samp1 proteinet har också visats orsaka denna sjukdom, vilket gav oss intresse att undersöka dessa mutationer i cellerna. Vi undersökte också om Samp1 hade en roll i differentiering av stamceller och nervceller.

35

8. Acknowledgements

Acknowledgments. First, I want to give a special thanks to all the people that read my thesis especially my opponent Philippe Collas and the committee members Maria Eriksson, Ann- Kristin Östlund Farrants, Anthony Wright and Daniel Daley.

Second, I want to thank my colleagues. I´m so happy to have you all in my life. This whole PhD-period would not have been the same without you.

To my supervisor, Einar Hallberg, thank you for letting me be a part of your research group and develop my scientific skills in your lab. I have learned so much under your supportive supervision. You always take time to find new ideas, discuss results and motivate us. You gave me the chance to independently develop while believing in me. In your research group I found people I will always consider my friends. The scientific experience, energy and kindness was a great part of our research group. Thank you all for sharing your knowledge and experience, and for all the great times we spent together. I hope we continue with our outside of work “lab meetings”. Veronica, I never meet a friend with the ability to convince me to do anything as easy as you do. I miss your kindness, laughter and good heart in the lab. Our trip to Switzerland is one of my favorite memories. Hakim, through all our struggles and hard work at the lab, we had really great teamwork together. I miss your calmness, industriousness and friendship at work. Balaje, thanks for your life energy and joy that you brought to the lab and your reassurance about the future. You are a truthful and caring person that I´m happy to call my friend. Elena, thank you for all our discussions about work and life. You are a kind and unique person that I´m happy I got to know in the last years. Thank you: Frida, you always impress me with your general knowledge (in grammar, birds, geography…). It was a real delight to work in a project with you, we really complemented each other. I´m sure that you will succeed in whatever you take on in the future. Mehedi, for keep trying, never giving up and being the best boule teammate. Good luck with the rest of your Ph.D. Ricardo, for your scientific expertise and always trying to help. You are a good teacher and I hope you keep inspiring your students.

Thanks to all my students and coauthors for your great contributions, research is teamwork and without you, the projects would not have been the same. Urška, thanks for changing your masters and coming back. For all the help, compliments and happiness. It was a real joy to getting to know you and I hope our paths cross again. Yohalie, thank you for bringing the most interesting but also the toughest project. We really learned to solve problems together. I wish you good luck with your Ph.D. and hope to follow the advancements. Thank you: Martin, for never giving up on the impossible method. Thank you for all the good times in the lab. Paulo, for your contributions and final success, that expanded our tools in the lab. Fausto, for your dedication and joyfulness in the lab.

To my old friends and colleagues at the neurochemistry corridor, thank you for making this a great work environment. For all the traditions, parties, barbecues, sports, dinners and lunches, we had together. I will miss you all and always remember this time with happiness. Anna, my “work-twin”, for all things we shared; office, birthday, adventures but mostly our talks. You always take the time to listen and being a really good friend. You are the best office-friend I could have wished for. Thank you: Niina, for your encouragement and support. Preeti, for your kindness and our talks. Kristina A, for your joy and energy. Maxime, for your stability and kindness. Aram, for your calm and thoughtfulness. Dan, for your strange jokes and skiing. Andreas, for your calm and joy. Ylva, for your professionalism and happiness. Luca, 36 for being nice and helpful. Helena, for your interest and care. Birgitta, for your pursuit of development and friendliness. Kristina W, for your kindness and compliments. Xin, for your support and friendliness. Jakob, for your organizing skill and keeping track of everything. Andrés, for your energy and amusingness. Carmine, for your optimism and energy. Jonas, for organizing many common activities and your kindness. Tönis, for being straightforward and all your initiatives. Jessica, for your charisma and friendliness. Moataz, for your dedication to learn and calmness. Henrik, for your IT help and encouragements. Ying, for your kindness and joy. Sylvia and Marie-Louise, for taking care of us. You kept the workplace in order and helped us with all the administrative tasks.

A special thanks, to all group leaders at Neurochemistry for your interesting research and care in us. You motivated us with your hard-work, we could always depend on you wanting our best.

Thanks to my new friends at the DBB corridor. Thank you: Mehwish, for your inspiration, enthusiasm and friendship. You brought joy to even the smallest things as hail, laying down in the grass or standing in the front on a boat. Theresa, for your energy, inspiration and involvement. Time always runs away when talking to you. I´m happy we became friends and I will always remember the bulldozer trip. Sania, for welcoming me to DBB. You were a good office-friend and I miss our “Swenglish” talks. Ingrid, for your involvement, optimism and enthusiasm. I hope we continue with all our activities. Rageia, for always cheering me up with our talks. Ane, for all your compliments and encouragement. James, for your energy and appetite. Alexandros, for always being cheerful and talkative. Daphne and Felix, for always being nice, you will be such great parents. Nir and Grant, for always being nice and helpful. Diana, for being straightforward and open. Dirk, for your traditions and stability. Andrea, for your kindness and optimism. Cecilia, for reminding me about our breaks and being a caring person. Matt, for being really nice and always fixing everything in record speed. Alexander, for all your help.

Last, I want to thank my friends and family for always believing in me and keeping my motivation up. I am so happy I have you all in my life. For always going along with all my events and crazy ideas. For making me feel smart and loved, I’m finally Dr.Cill. Thank you for supporting me. To my handsome husband William, for your unconditional love and support. From the first meeting at the jumping castle to all the adventures and life stages we have experienced together, we have always been the perfect team. You balance me and make me grow. You turn every situation into something that makes us even closer. “When the tears come streaming down my face, you will guide me home”. You are my matching puzzle piece and I can´t imagine my life without you. To Alexander and Alice, the love of my life. Alexander, you are sociable, remarkable and adventurous. Alice, you are considerate, independent and cute. My child, just like the nucleus, you are the center of my cell. From your first moment to my last, I will be there loving every memory.

37

9. References

Ahmad, K., Henikoff, S., 2002. The Histone Variant H3.3 Marks Active Chromatin by Replication-Independent Nucleosome Assembly. Mol. Cell 9, 1191–1200. https://doi.org/10.1016/S1097-2765(02)00542-7 Alva, V., Ammelburg, M., Söding, J., Lupas, A.N., 2007. On the origin of the histone fold. BMC Struct. Biol. 7, 17. https://doi.org/10.1186/1472-6807-7-17 Amendola, M., van Steensel, B., 2014. Mechanisms and dynamics of nuclear lamina–genome interactions. Curr. Opin. Cell Biol. 28, 61–68. https://doi.org/10.1016/j.ceb.2014.03.003 Barr, A.R., Gergely, F., 2007. Aurora-A: the maker and breaker of spindle poles. J. Cell Sci. 120, 2987–2996. https://doi.org/10.1242/jcs.013136 Barrales, R.R., Forn, M., Georgescu, P.R., Sarkadi, Z., Braun, S., 2016. Control of heterochromatin localization and silencing by the nuclear membrane protein Lem2. Genes Dev. 30, 133–148. https://doi.org/10.1101/gad.271288.115 Batrakou, D.G., de Las Heras, J.I., Czapiewski, R., Mouras, R., Schirmer, E.C., 2015. TMEM120A and B: Nuclear Envelope Transmembrane Proteins Important for Adipocyte Differentiation. PloS One 10, e0127712. https://doi.org/10.1371/journal.pone.0127712 Bergqvist, C., 2016. The role of nuclear membrane proteins in differentiation and chromatin organization (Licentiate thesis). SU, ISBN 978-91-7649-632-9. Berk, J.M., Tifft, K.E., Wilson, K.L., 2013. The nuclear envelope LEM-domain protein emerin. Nucleus 4, 298–314. https://doi.org/10.4161/nucl.25751 Bhaskara, S., Knutson, S.K., Jiang, G., Chandrasekharan, M.B., Wilson, A.J., Zheng, S., Yenamandra, A., Locke, K., Yuan, J., Bonine-Summers, A.R., Wells, C.E., Kaiser, J.F., Washington, M.K., Zhao, Z., Wagner, F.F., Sun, Z.-W., Xia, F., Holson, E.B., Khabele, D., Hiebert, S.W., 2010. Hdac3 is essential for the maintenance of chromatin structure and genome stability. Cancer Cell 18, 436–447. https://doi.org/10.1016/j.ccr.2010.10.022 Bickmore, W.A., van Steensel, B., 2013. Genome architecture: domain organization of interphase chromosomes. Cell 152, 1270–1284. https://doi.org/10.1016/j.cell.2013.02.001 Bione, S., Maestrini, E., Rivella, S., Mancini, M., Regis, S., Romeo, G., Toniolo, D., 1994. Identification of a novel X-linked gene responsible for Emery-Dreifuss muscular dystrophy. Nat. Genet. 8, 323–327. https://doi.org/10.1038/ng1294-323 Black, B.E., Foltz, D.R., Chakravarthy, S., Luger, K., Woods, V.L., Cleveland, D.W., 2004. Structural determinants for generating centromeric chromatin. Nature 430, 578– 582. https://doi.org/10.1038/nature02766 Bonne, G., Leturcq, F., Ben Yaou, R., 1993. Emery-Dreifuss Muscular Dystrophy, in: Adam, M.P., Ardinger, H.H., Pagon, R.A., Wallace, S.E., Bean, L.J., Stephens, K., Amemiya, A. (Eds.), GeneReviews®. University of Washington, Seattle, Seattle (WA). Borrego-Pinto, J., Jegou, T., Osorio, D.S., Auradé, F., Gorjánácz, M., Koch, B., Mattaj, I.W., Gomes, E.R., 2012. Samp1 is a component of TAN lines and is required for nuclear movement. J. Cell Sci. 125, 1099–1105. https://doi.org/10.1242/jcs.087049 Bos, J.L., Rehmann, H., Wittinghofer, A., 2007. GEFs and GAPs: Critical Elements in the Control of Small G Proteins. Cell 129, 865–877. https://doi.org/10.1016/j.cell.2007.05.018

38

Bouzid, T., Kim, E., Riehl, B.D., Esfahani, A.M., Rosenbohm, J., Yang, R., Duan, B., Lim, J.Y., 2019. The LINC complex, mechanotransduction, and mesenchymal stem cell function and fate. J. Biol. Eng. 13. https://doi.org/10.1186/s13036-019-0197-9 Broers, J.L., Machiels, B.M., Kuijpers, H.J., Smedts, F., van den Kieboom, R., Raymond, Y., Ramaekers, F.C., 1997. A- and B-type lamins are differentially expressed in normal human tissues. Histochem. Cell Biol. 107, 505–517. https://doi.org/10.1007/s004180050138 Buch, C., Lindberg, R., Figueroa, R., Gudise, S., Onischenko, E., Hallberg, E., 2009. An integral protein of the inner nuclear membrane localizes to the mitotic spindle in mammalian cells. J. Cell Sci. 122, 2100–2107. https://doi.org/10.1242/jcs.047373 Burke, B., 2019. Chain reaction: LINC complexes and nuclear positioning. F1000Research 8. https://doi.org/10.12688/f1000research.16877.1 Burke, B., Stewart, C.L., 2014. Functional architecture of the cell’s nucleus in development, aging, and disease. Curr. Top. Dev. Biol. 109, 1–52. https://doi.org/10.1016/B978-0- 12-397920-9.00006-8 Burke, B., Stewart, C.L., 2002. Life at the edge: the nuclear envelope and human disease. Nat. Rev. Mol. Cell Biol. 3, 575–585. https://doi.org/10.1038/nrm879 Butterworth, P.J., 2005. Lehninger: principles of biochemistry (4th edn) D. L. Nelson and M. C. Cox, W. H. Freeman & Co., New York, 1119 pp (plus 17 pp glossary), ISBN 0-7167- 4339-6 (2004). Cell Biochem. Funct. 23, 293–294. https://doi.org/10.1002/cbf.1216 Cabal, G.G., Genovesio, A., Rodriguez-Navarro, S., Zimmer, C., Gadal, O., Lesne, A., Buc, H., Feuerbach-Fournier, F., Olivo-Marin, J.-C., Hurt, E.C., Nehrbass, U., 2006. SAGA interacting factors confine sub-diffusion of transcribed genes to the nuclear envelope. Nature 441, 770–773. https://doi.org/10.1038/nature04752 Cestan R, Lejonne NI, 1902. Une myopathie avec retractions familiales. Chal, J., Pourquié, O., 2017. Making muscle: skeletal myogenesis in vivo and in vitro. Development 144, 2104–2122. https://doi.org/10.1242/dev.151035 Chambers, I., Tomlinson, S.R., 2009. The transcriptional foundation of pluripotency. Dev. Camb. Engl. 136, 2311–2322. https://doi.org/10.1242/dev.024398 Chang, W., Antoku, S., Östlund, C., Worman, H.J., Gundersen, G.G., 2015a. Linker of nucleoskeleton and cytoskeleton (LINC) complex-mediated actin-dependent nuclear positioning orients centrosomes in migrating myoblasts. Nucleus 6, 77–88. https://doi.org/10.1080/19491034.2015.1004947 Chang, W., Worman, H.J., Gundersen, G.G., 2015b. Accessorizing and anchoring the LINC complex for multifunctionality. J. Cell Biol. 208, 11–22. https://doi.org/10.1083/jcb.201409047 Chen, L., Jiang, F., Qiao, Y., Li, H., Wei, Z., Huang, T., Lan, J., Xia, Y., Li, J., 2018. Nucleoskeletal stiffness regulates stem cell migration and differentiation through lamin A/C. J. Cell. Physiol. 233, 5112–5118. https://doi.org/10.1002/jcp.26336 Chen, R.-H., 2012. Nuclear Envelope Assembly and Disassembly During the Cell Cycle, in: ELS. American Cancer Society. https://doi.org/10.1002/9780470015902.a0022532 Chi, Y.-H., Haller, K., Peloponese, J.-M., Jeang, K.-T., 2007. Histone acetyltransferase hALP and nuclear membrane protein hsSUN1 function in de-condensation of mitotic chromosomes. J. Biol. Chem. 282, 27447–27458. https://doi.org/10.1074/jbc.M703098200 Chin, M.H., Mason, M.J., Xie, W., Volinia, S., Singer, M., Peterson, C., Ambartsumyan, G., Aimiuwu, O., Richter, L., Zhang, J., Khvorostov, I., Ott, V., Grunstein, M., Lavon, N.,

39

Benvenisty, N., Croce, C.M., Clark, A.T., Baxter, T., Pyle, A.D., Teitell, M.A., Pelegrini, M., Plath, K., Lowry, W.E., 2009. Induced pluripotent stem cells and embryonic stem cells are distinguished by gene expression signatures. Cell Stem Cell 5, 111–123. https://doi.org/10.1016/j.stem.2009.06.008 Chook, Y., Blobel, G., 2001. Karyopherins and nuclear import. Curr. Opin. Struct. Biol. 11, 703–715. https://doi.org/10.1016/S0959-440X(01)00264-0 Clarke, P.R., Zhang, C., 2008. Spatial and temporal coordination of mitosis by Ran GTPase. Nat. Rev. Mol. Cell Biol. 9, 464–477. https://doi.org/10.1038/nrm2410 Clowney, E.J., LeGros, M.A., Mosley, C.P., Clowney, F.G., Markenskoff-Papadimitriou, E.C., Myllys, M., Barnea, G., Larabell, C.A., Lomvardas, S., 2012. Nuclear Aggregation of Olfactory Receptor Genes Governs Their Monogenic Expression. Cell 151, 724–737. https://doi.org/10.1016/j.cell.2012.09.043 Coffinier, C., Jung, H.-J., Nobumori, C., Chang, S., Tu, Y., Barnes, R.H., Yoshinaga, Y., de Jong, P.J., Vergnes, L., Reue, K., Fong, L.G., Young, S.G., 2011. Deficiencies in lamin B1 and lamin B2 cause neurodevelopmental defects and distinct nuclear shape abnormalities in neurons. Mol. Biol. Cell 22, 4683–4693. https://doi.org/10.1091/mbc.E11-06-0504 Constantinescu, D., Gray, H.L., Sammak, P.J., Schatten, G.P., Csoka, A.B., 2006. Lamin A/C expression is a marker of mouse and human embryonic stem cell differentiation. Stem Cells Dayt. Ohio 24, 177–185. https://doi.org/10.1634/stemcells.2004-0159 Cremer, T., Cremer, M., 2010. Chromosome Territories. Cold Spring Harb. Perspect. Biol. 2. https://doi.org/10.1101/cshperspect.a003889 Crisp, M., Liu, Q., Roux, K., Rattner, J.B., Shanahan, C., Burke, B., Stahl, P.D., Hodzic, D., 2006. Coupling of the nucleus and cytoplasm: role of the LINC complex. J. Cell Biol. 172, 41–53. https://doi.org/10.1083/jcb.200509124 Croft, J.A., Bridger, J.M., Boyle, S., Perry, P., Teague, P., Bickmore, W.A., 1999. Differences in the localization and morphology of chromosomes in the human nucleus. J. Cell Biol. 145, 1119–1131. https://doi.org/10.1083/jcb.145.6.1119 Dauer, W.T., Worman, H.J., 2009. The nuclear envelope as a signaling node in development and disease. Dev. Cell 17, 626–638. https://doi.org/10.1016/j.devcel.2009.10.016 Davies, H.G., 1967. Fine Structure of Heterochromatin in Certain Cell Nuclei. Nature 214, 208–210. https://doi.org/10.1038/214208a0 De Sandre-Giovannoli, A., Bernard, R., Cau, P., Navarro, C., Amiel, J., Boccaccio, I., Lyonnet, S., Stewart, C.L., Munnich, A., Le Merrer, M., Lévy, N., 2003. Lamin a truncation in Hutchinson-Gilford progeria. Science 300, 2055. https://doi.org/10.1126/science.1084125 Dechat, T., Pfleghaar, K., Sengupta, K., Shimi, T., Shumaker, D.K., Solimando, L., Goldman, R.D., 2008. Nuclear lamins: major factors in the structural organization and function of the nucleus and chromatin. Genes Dev. 22, 832–853. https://doi.org/10.1101/gad.1652708 Delbarre, E., Jacobsen, B.M., Reiner, A.H., Sørensen, A.L., Küntziger, T., Collas, P., 2010. Chromatin Environment of Histone Variant H3.3 Revealed by Quantitative Imaging and Genome-scale Chromatin and DNA Immunoprecipitation. Mol. Biol. Cell 21, 1872–1884. https://doi.org/10.1091/mbc.E09-09-0839 Demmerle, J., Koch, A.J., Holaska, J.M., 2012. The Nuclear Envelope Protein Emerin Binds Directly to Histone Deacetylase 3 (HDAC3) and Activates HDAC3 Activity. J. Biol. Chem. 287, 22080–22088. https://doi.org/10.1074/jbc.M111.325308

40

Dixon, J.R., Jung, I., Selvaraj, S., Shen, Y., Antosiewicz-Bourget, J.E., Lee, A.Y., Ye, Z., Kim, A., Rajagopal, N., Xie, W., Diao, Y., Liang, J., Zhao, H., Lobanenkov, V.V., Ecker, J.R., Thomson, J.A., Ren, B., 2015. Chromatin architecture reorganization during stem cell differentiation. Nature 518, 331–336. https://doi.org/10.1038/nature14222 Eriksson, M., Brown, W.T., Gordon, L.B., Glynn, M.W., Singer, J., Scott, L., Erdos, M.R., Robbins, C.M., Moses, T.Y., Berglund, P., Dutra, A., Pak, E., Durkin, S., Csoka, A.B., Boehnke, M., Glover, T.W., Collins, F.S., 2003. Recurrent de novo point mutations in lamin A cause Hutchinson–Gilford progeria syndrome. Nature 423, 293–298. https://doi.org/10.1038/nature01629 Fahrenkrog, B., Aebi, U., 2003. The nuclear pore complex: Nucleocytoplasmic transport and beyond. Nat. Rev. Mol. Cell Biol. 4, 757–66. https://doi.org/10.1038/nrm1230 Figueroa, R., Gudise, S., Larsson, V., Hallberg, E., 2010. A transmembrane inner nuclear membrane protein in the mitotic spindle. Nucleus 1, 249–253. https://doi.org/10.4161/nucl.11740 Fiorotto, M., 2012. The making of a muscle. The biochemist 34, 4–11. Fišerová, J., Efenberková, M., Sieger, T., Maninová, M., Uhlířová, J., Hozák, P., 2017. Chromatin organization at the nuclear periphery as revealed by image analysis of structured illumination microscopy data. J. Cell Sci. 130, 2066–2077. https://doi.org/10.1242/jcs.198424 Fong, L.G., Ng, J.K., Lammerding, J., Vickers, T.A., Meta, M., Coté, N., Gavino, B., Qiao, X., Chang, S.Y., Young, S.R., Yang, S.H., Stewart, C.L., Lee, R.T., Bennett, C.F., Bergo, M.O., Young, S.G., 2006. Prelamin A and lamin A appear to be dispensable in the nuclear lamina. J. Clin. Invest. 116, 743–752. https://doi.org/10.1172/JCI27125 Forster, J.I., Köglsberger, S., Trefois, C., Boyd, O., Baumuratov, A.S., Buck, L., Balling, R., Antony, P.M.A., 2016. Characterization of Differentiated SH-SY5Y as Neuronal Screening Model Reveals Increased Oxidative Vulnerability. J. Biomol. Screen. 21, 496–509. https://doi.org/10.1177/1087057115625190 Francastel, C., Magis, W., Groudine, M., 2001. Nuclear relocation of a transactivator subunit precedes target gene activation. Proc. Natl. Acad. Sci. U. S. A. 98, 12120–12125. https://doi.org/10.1073/pnas.211444898 Gaspar-Maia, A., Alajem, A., Meshorer, E., Ramalho-Santos, M., 2011. Open chromatin in pluripotency and reprogramming. Nat. Rev. Mol. Cell Biol. 12, 36–47. https://doi.org/10.1038/nrm3036 Goshima, G., Mayer, M., Zhang, N., Stuurman, N., Vale, R.D., 2008. Augmin: a protein complex required for centrosome-independent microtubule generation within the spindle. J. Cell Biol. 181, 421–429. https://doi.org/10.1083/jcb.200711053 Gruenbaum, Y., Foisner, R., 2015. Lamins: nuclear proteins with fundamental functions in nuclear mechanics and genome regulation. Annu. Rev. Biochem. 84, 131–164. https://doi.org/10.1146/annurev-biochem-060614-034115 Gudise, S., Figueroa, R.A., Lindberg, R., Larsson, V., Hallberg, E., 2011. Samp1 is functionally associated with the LINC complex and A-type lamina networks. J. Cell Sci. 124, 2077– 2085. https://doi.org/10.1242/jcs.078923 Guerreiro, I., Kind, J., 2019. Spatial chromatin organization and gene regulation at the nuclear lamina. Curr. Opin. Genet. Dev. 55, 19–25. https://doi.org/10.1016/j.gde.2019.04.008 Guilluy, C., Osborne, L.D., Van Landeghem, L., Sharek, L., Superfine, R., Garcia-Mata, R., Burridge, K., 2014. Isolated nuclei adapt to force and reveal a mechanotransduction

41

pathway in the nucleus. Nat. Cell Biol. 16, 376–381. https://doi.org/10.1038/ncb2927 Hah, J., Kim, D.-H., 2019. Deciphering Nuclear Mechanobiology in . Cells 8. https://doi.org/10.3390/cells8030231 Harada, T., Swift, J., Irianto, J., Shin, J.-W., Spinler, K.R., Athirasala, A., Diegmiller, R., Dingal, P.C.D.P., Ivanovska, I.L., Discher, D.E., 2014. Nuclear lamin stiffness is a barrier to 3D migration, but softness can limit survival. J. Cell Biol. 204, 669–682. https://doi.org/10.1083/jcb.201308029 Harr, J.C., Gonzalez-Sandoval, A., Gasser, S.M., 2016. Histones and histone modifications in perinuclear chromatin anchoring: from yeast to man. EMBO Rep. 17, 139–155. https://doi.org/10.15252/embr.201541809 Heller, S.A., Shih, R., Kalra, R., Kang, P.B., 2020. Emery‐Dreifuss muscular dystrophy. Muscle Nerve 61, 436–448. https://doi.org/10.1002/mus.26782 Hetzer, M.W., 2010. The nuclear envelope. Cold Spring Harb. Perspect. Biol. 2, a000539. https://doi.org/10.1101/cshperspect.a000539 Ho, R., Hegele, R.A., 2019. Complex effects of laminopathy mutations on nuclear structure and function. Clin. Genet. 95, 199–209. https://doi.org/10.1111/cge.13455 Holaska, J.M., 2008. Emerin and the nuclear lamina in muscle and cardiac disease. Circ. Res. 103, 16–23. https://doi.org/10.1161/CIRCRESAHA.108.172197 Houben, F., Ramaekers, F.C.S., Snoeckx, L.H.E.H., Broers, J.L.V., 2007. Role of nuclear lamina- cytoskeleton interactions in the maintenance of cellular strength. Biochim. Biophys. Acta BBA - Mol. Cell Res., Integrated approaches to cytoskeleton research 1773, 675–686. https://doi.org/10.1016/j.bbamcr.2006.09.018 Huber, M.D., Guan, T., Gerace, L., 2009. Overlapping functions of nuclear envelope proteins NET25 (Lem2) and emerin in regulation of extracellular signal-regulated kinase signaling in myoblast differentiation. Mol. Cell. Biol. 29, 5718–5728. https://doi.org/10.1128/MCB.00270-09 Ibarra, A., Hetzer, M.W., 2015. Nuclear pore proteins and the control of genome functions. Genes Dev. 29, 337–349. https://doi.org/10.1101/gad.256495.114 Ivanauskiene, K., Delbarre, E., McGhie, J.D., Küntziger, T., Wong, L.H., Collas, P., 2014. The PML-associated protein DEK regulates the balance of H3.3 loading on chromatin and is important for telomere integrity. Genome Res. 24, 1584–1594. https://doi.org/10.1101/gr.173831.114 Jafferali, M.H., Figueroa, R.A., Hasan, M., Hallberg, E., 2017. Spindle associated membrane protein 1 (Samp1) is required for the differentiation of muscle cells. Sci. Rep. 7, 16655. https://doi.org/10.1038/s41598-017-16746-y Jafferali, M.H., Vijayaraghavan, B., Figueroa, R.A., Crafoord, E., Gudise, S., Larsson, V.J., Hallberg, E., 2014. MCLIP, an effective method to detect interactions of transmembrane proteins of the nuclear envelope in live cells. Biochim. Biophys. Acta 1838, 2399–2403. https://doi.org/10.1016/j.bbamem.2014.06.008 Jahed, Z., Soheilypour, M., Peyro, M., Mofrad, M.R.K., 2016. The LINC and NPC relationship - it’s complicated! J. Cell Sci. 129, 3219–3229. https://doi.org/10.1242/jcs.184184 Janin, A., Bauer, D., Ratti, F., Millat, G., Méjat, A., 2017. Nuclear envelopathies: a complex LINC between nuclear envelope and pathology. Orphanet J. Rare Dis. 12, 147. https://doi.org/10.1186/s13023-017-0698-x Jenuwein, T., Allis, C.D., 2001. Translating the histone code. Science 293, 1074–1080. https://doi.org/10.1126/science.1063127

42

Jerabek-Willemsen, M., Wienken, C.J., Braun, D., Baaske, P., Duhr, S., 2011. Molecular Interaction Studies Using Microscale Thermophoresis. Assay Drug Dev. Technol. 9, 342–353. https://doi.org/10.1089/adt.2011.0380 Ji, P., Murata-Hori, M., Lodish, H.F., 2011. Formation of mammalian erythrocytes: Chromatin condensation and enucleation. Trends Cell Biol. 21, 409–415. https://doi.org/10.1016/j.tcb.2011.04.003 Kanda, T., Sullivan, K.F., Wahl, G.M., 1998. Histone-GFP fusion protein enables sensitive analysis of chromosome dynamics in living mammalian cells. Curr. Biol. CB 8, 377– 385. https://doi.org/10.1016/s0960-9822(98)70156-3 Katta, S.S., Smoyer, C.J., Jaspersen, S.L., 2014. Destination: inner nuclear membrane. Trends Cell Biol. 24, 221–229. https://doi.org/10.1016/j.tcb.2013.10.006 Kim, Y.J., Cecchini, K.R., Kim, T.H., 2011. Conserved, developmentally regulated mechanism couples chromosomal looping and heterochromatin barrier activity at the homeobox gene A locus. Proc. Natl. Acad. Sci. U. S. A. 108, 7391–7396. https://doi.org/10.1073/pnas.1018279108 Kimura, H., Cook, P.R., 2001. Kinetics of core histones in living human cells: little exchange of H3 and H4 and some rapid exchange of H2B. J. Cell Biol. 153, 1341–1353. https://doi.org/10.1083/jcb.153.7.1341 King, M.C., Drivas, T.G., Blobel, G., 2008. A network of nuclear envelope membrane proteins linking centromeres to microtubules. Cell 134, 427–438. https://doi.org/10.1016/j.cell.2008.06.022 Kite, G.L., 1913. The Relative Permeability of the Surface and Interior Portions of the Cytoplasm of and Plant Cells. (A Preliminary Paper). Biol. Bull. 25, 1–7. https://doi.org/10.2307/1536080 Koch, A.J., Holaska, J.M., 2014. Emerin in health and disease. Semin. Cell Dev. Biol. 29, 95– 106. https://doi.org/10.1016/j.semcdb.2013.12.008 Korfali, N., Wilkie, G.S., Swanson, S.K., Srsen, V., Batrakou, D.G., Fairley, E.A.L., Malik, P., Zuleger, N., Goncharevich, A., de Las Heras, J., Kelly, D.A., Kerr, A.R.W., Florens, L., Schirmer, E.C., 2010. The leukocyte nuclear envelope proteome varies with cell activation and contains novel transmembrane proteins that affect genome architecture. Mol. Cell. Proteomics MCP 9, 2571–2585. https://doi.org/10.1074/mcp.M110.002915 Korfali, N., Wilkie, G.S., Swanson, S.K., Srsen, V., de las Heras, J., Batrakou, D.G., Malik, P., Zuleger, N., Kerr, A.R.W., Florens, L., Schirmer, E.C., 2012. The nuclear envelope proteome differs notably between tissues. Nucleus 3, 552–564. https://doi.org/10.4161/nucl.22257 Kornberg, R.D., 1974. Chromatin structure: a repeating unit of histones and DNA. Science 184, 868–871. https://doi.org/10.1126/science.184.4139.868 Kouzarides, T., 2007. Chromatin modifications and their function. Cell 128, 693–705. https://doi.org/10.1016/j.cell.2007.02.005 Kouzarides, T., 2000. Acetylation: a regulatory modification to rival ? EMBO J. 19, 1176–1179. https://doi.org/10.1093/emboj/19.6.1176 Kumaran, R.I., Spector, D.L., 2008. A genetic locus targeted to the nuclear periphery in living cells maintains its transcriptional competence. J. Cell Biol. 180, 51–65. https://doi.org/10.1083/jcb.200706060

43

Larsson, V.J., Jafferali, M.H., Vijayaraghavan, B., Figueroa, R.A., Hallberg, E., 2018. Mitotic spindle assembly and γ-tubulin localisation depend on the integral nuclear membrane protein Samp1. J. Cell Sci. 131. https://doi.org/10.1242/jcs.211664 Lattanzi, G., Ortolani, M., Columbaro, M., Prencipe, S., Mattioli, E., Lanzarini, C., Maraldi, N.M., Cenni, V., Garagnani, P., Salvioli, S., Storci, G., Bonafè, M., Capanni, C., Franceschi, C., 2014. Lamins are rapamycin targets that impact human longevity: a study in centenarians. J. Cell Sci. 127, 147–157. https://doi.org/10.1242/jcs.133983 Lieberman-Aiden, E., van Berkum, N.L., Williams, L., Imakaev, M., Ragoczy, T., Telling, A., Amit, I., Lajoie, B.R., Sabo, P.J., Dorschner, M.O., Sandstrom, R., Bernstein, B., Bender, M.A., Groudine, M., Gnirke, A., Stamatoyannopoulos, J., Mirny, L.A., Lander, E.S., Dekker, J., 2009. Comprehensive mapping of long range interactions reveals folding principles of the human genome. Science 326, 289–293. https://doi.org/10.1126/science.1181369 Lin, C.-J., Conti, M., Ramalho-Santos, M., 2013. Histone variant H3.3 maintains a decondensed chromatin state essential for mouse preimplantation development. Development 140, 3624–3634. https://doi.org/10.1242/dev.095513 Liu, S., Pellman, D., 2020. The coordination of nuclear envelope assembly and chromosome segregation in metazoans. Nucleus 11, 35–52. https://doi.org/10.1080/19491034.2020.1742064 Loyola, A., Almouzni, G., 2007. Marking histone H3 variants: How, when and why? Trends Biochem. Sci. 32, 425–433. https://doi.org/10.1016/j.tibs.2007.08.004 Lui, K., Huang, Y., 2009. RanGTPase: A Key Regulator of Nucleocytoplasmic Trafficking. Mol. Cell. Pharmacol. 1, 148–156. Machiels, B.M., Zorenc, A.H., Endert, J.M., Kuijpers, H.J., van Eys, G.J., Ramaekers, F.C., Broers, J.L., 1996. An alternative splicing product of the lamin A/C gene lacks exon 10. J. Biol. Chem. 271, 9249–9253. https://doi.org/10.1074/jbc.271.16.9249 Madej-Pilarczyk, A., 2018. Clinical aspects of Emery-Dreifuss muscular dystrophy. Nucleus 9, 268–274. https://doi.org/10.1080/19491034.2018.1462635 Maimon, T., Elad, N., Dahan, I., Medalia, O., 2012. The Human Nuclear Pore Complex as Revealed by Cryo-Electron Tomography. Structure 20, 998–1006. https://doi.org/10.1016/j.str.2012.03.025 Maresca, G., Natoli, M., Nardella, M., Arisi, I., Trisciuoglio, D., Desideri, M., Brandi, R., D’Aguanno, S., Nicotra, M.R., D’Onofrio, M., Urbani, A., Natali, P.G., Del Bufalo, D., Felsani, A., D’Agnano, I., 2012. LMNA knock-down affects differentiation and progression of human neuroblastoma cells. PloS One 7, e45513. https://doi.org/10.1371/journal.pone.0045513 Margalit, A., Brachner, A., Gotzmann, J., Foisner, R., Gruenbaum, Y., 2007. Barrier-to- autointegration factor – a BAFfling little protein. Trends Cell Biol. 17, 202–208. https://doi.org/10.1016/j.tcb.2007.02.004 Masters, J.R., 2002. HeLa cells 50 years on: the good, the bad and the ugly. Nat. Rev. Cancer 2, 315–319. https://doi.org/10.1038/nrc775 Mattioli, E., Columbaro, M., Capanni, C., Maraldi, N.M., Cenni, V., Scotlandi, K., Marino, M.T., Merlini, L., Squarzoni, S., Lattanzi, G., 2011. Prelamin A-mediated recruitment of SUN1 to the nuclear envelope directs nuclear positioning in human muscle. Cell Death Differ. 18, 1305–1315. https://doi.org/10.1038/cdd.2010.183

44

Mattioli, E., Columbaro, M., Jafferali, M.H., Schena, E., Hallberg, E., Lattanzi, G., 2018. Samp1 Mislocalization in Emery-Dreifuss Muscular Dystrophy. Cells 7. https://doi.org/10.3390/cells7100170 Meinke, P., Kerr, A.R.W., Czapiewski, R., Heras, J.I. de las, Dixon, C.R., Harris, E., Kölbel, H., Muntoni, F., Schara, U., Straub, V., Schoser, B., Wehnert, M., Schirmer, E.C., 2020. A multistage sequencing strategy pinpoints novel candidate alleles for Emery-Dreifuss muscular dystrophy and supports gene misregulation as its pathomechanism. EBioMedicine 51. https://doi.org/10.1016/j.ebiom.2019.11.048 Melchior, F., 2001. Ran GTPase cycle: One mechanism — two functions. Curr. Biol. 11, R257–R260. https://doi.org/10.1016/S0960-9822(01)00132-4 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., Brooks, B.P., Gerber, L.H., Turner, M.L., Domingo, D.L., Hart, T.C., Graf, J., Reynolds, J.C., Gropman, A., Yanovski, J.A., Gerhard-Herman, M., Collins, F.S., Nabel, E.G., Cannon, R.O., Gahl, W.A., Introne, W.J., 2008. Phenotype and course of Hutchinson-Gilford progeria syndrome. N. Engl. J. Med. 358, 592–604. https://doi.org/10.1056/NEJMoa0706898 Meunier, S., Vernos, I., 2012. Microtubule assembly during mitosis - from distinct origins to distinct functions? J. Cell Sci. 125, 2805–2814. https://doi.org/10.1242/jcs.092429 Mills, J.C., Wang, S., Erecińska, M., Pittman, R.N., 1995. Use of cultured neurons and neuronal cell lines to study morphological, biochemical, and molecular changes occurring in cell death. Methods Cell Biol. 46, 217–242. https://doi.org/10.1016/s0091-679x(08)61931-7 Mohn, F., Weber, M., Rebhan, M., Roloff, T.C., Richter, J., Stadler, M.B., Bibel, M., Schübeler, D., 2008. Lineage-specific polycomb targets and de novo DNA methylation define restriction and potential of neuronal progenitors. Mol. Cell 30, 755–766. https://doi.org/10.1016/j.molcel.2008.05.007 Morris, G.E., 2001. The role of the nuclear envelope in Emery-Dreifuss muscular dystrophy. Trends Mol. Med. 7, 572–577. https://doi.org/10.1016/s1471-4914(01)02128-1 Muchir, A., Shan, J., Bonne, G., Lehnart, S.E., Worman, H.J., 2009a. Inhibition of extracellular signal-regulated kinase signaling to prevent cardiomyopathy caused by mutation in the gene encoding A-type lamins. Hum. Mol. Genet. 18, 241–247. https://doi.org/10.1093/hmg/ddn343 Muchir, A., Wu, W., Worman, H.J., 2009b. Reduced expression of A-type lamins and emerin activates extracellular signal-regulated kinase in cultured cells. Biochim. Biophys. Acta 1792, 75–81. https://doi.org/10.1016/j.bbadis.2008.10.012 Mudumbi, K.C., Czapiewski, R., Ruba, A., Junod, S.L., Li, Y., Luo, W., Ngo, C., Ospina, V., Schirmer, E.C., Yang, W., 2020. Nucleoplasmic signals promote directed transmembrane protein import simultaneously via multiple channels of nuclear pores. Nat. Commun. 11. https://doi.org/10.1038/s41467-020-16033-x Musich, P.R., Zou, Y., 2009. Genomic Instability and DNA Damage Responses in Progeria Arising from Defective Maturation of Prelamin A. Aging 1, 28–37. Neumann, B., Walter, T., Hériché, J.-K., Bulkescher, J., Erfle, H., Conrad, C., Rogers, P., Poser, I., Held, M., Liebel, U., Cetin, C., Sieckmann, F., Pau, G., Kabbe, R., Wünsche, A., Satagopam, V., Schmitz, M.H.A., Chapuis, C., Gerlich, D.W., Schneider, R., Eils, R., Huber, W., Peters, J.-M., Hyman, A.A., Durbin, R., Pepperkok, R., Ellenberg, J., 2010. Phenotypic profiling of the human genome by time-lapse microscopy reveals genes. Nature 464, 721–727. https://doi.org/10.1038/nature08869

45

Nikolakaki, E., Mylonis, I., Giannakouros, T., 2017. Lamin B Receptor: Interplay between Structure, Function and Localization. Cells 6. https://doi.org/10.3390/cells6030028 Nishitani, H., Ohtsubo, M., Yamashita, K., Iida, H., Pines, J., Yasudo, H., Shibata, Y., Hunter, T., Nishimoto, T., 1991. Loss of RCC1, a nuclear DNA-binding protein, uncouples the completion of DNA replication from the activation of cdc2 and mitosis. EMBO J. 10, 1555–1564. https://doi.org/10.1002/j.1460- 2075.1991.tb07675.x Nmezi, B., Xu, J., Fu, R., Armiger, T.J., Rodriguez-Bey, G., Powell, J.S., Ma, H., Sullivan, M., Tu, Y., Chen, N.Y., Young, S.G., Stolz, D.B., Dahl, K.N., Liu, Y., Padiath, Q.S., 2019. Concentric organization of A- and B-type lamins predicts their distinct roles in the spatial organization and stability of the nuclear lamina. Proc. Natl. Acad. Sci. 116, 4307–4315. https://doi.org/10.1073/pnas.1810070116 Norwood, F.L.M., Harling, C., Chinnery, P.F., Eagle, M., Bushby, K., Straub, V., 2009. Prevalence of genetic muscle disease in Northern England: in-depth analysis of a muscle clinic population. Brain J. Neurol. 132, 3175–3186. https://doi.org/10.1093/brain/awp236 Parada, L.A., McQueen, P.G., Misteli, T., 2004. Tissue-specific spatial organization of . Genome Biol. 5, R44. https://doi.org/10.1186/gb-2004-5-7-r44 Parry, A.J., Narita, M., 2016. Old cells, new tricks: chromatin structure in senescence. Mamm. Genome 27, 320–331. https://doi.org/10.1007/s00335-016-9628-9 Peric-Hupkes, D., Meuleman, W., Pagie, L., Bruggeman, S.W.M., Solovei, I., Brugman, W., Gräf, S., Flicek, P., Kerkhoven, R.M., van Lohuizen, M., Reinders, M., Wessels, L., van Steensel, B., 2010. Molecular Maps of the Reorganization of Genome-Nuclear Lamina Interactions during Differentiation. Mol. Cell 38, 603–613. https://doi.org/10.1016/j.molcel.2010.03.016 Pittman, R.N., Wang, S., DiBenedetto, A.J., Mills, J.C., 1993. A system for characterizing cellular and molecular events in programmed neuronal cell death. J. Neurosci. Off. J. Soc. Neurosci. 13, 3669–3680. Ponten, J., Saksela, E., 1967. Two established in vitro cell lines from human mesenchymal tumours. Int. J. Cancer 2, 434–447. https://doi.org/10.1002/ijc.2910020505 Puckelwartz, M.J., Depreux, F.F., McNally, E.M., 2011. Gene expression, chromosome position and lamin A/C mutations. Nucleus 2, 162–167. https://doi.org/10.4161/nucl.2.3.16003 Ricci, M.A., Manzo, C., García-Parajo, M.F., Lakadamyali, M., Cosma, M.P., 2015. Chromatin fibers are formed by heterogeneous groups of nucleosomes in vivo. Cell 160, 1145– 1158. https://doi.org/10.1016/j.cell.2015.01.054 Ritland Politz, J.C., Scalzo, D., Groudine, M., 2016. The redundancy of the mammalian heterochromatic compartment. Curr. Opin. Genet. Dev. 37, 1–8. https://doi.org/10.1016/j.gde.2015.10.007 Robson, M.I., de Las Heras, J.I., Czapiewski, R., Lê Thành, P., Booth, D.G., Kelly, D.A., Webb, S., Kerr, A.R.W., Schirmer, E.C., 2016. Tissue-Specific Gene Repositioning by Muscle Nuclear Membrane Proteins Enhances Repression of Critical Developmental Genes during Myogenesis. Mol. Cell 62, 834–847. https://doi.org/10.1016/j.molcel.2016.04.035 Salpingidou, G., Smertenko, A., Hausmanowa-Petrucewicz, I., Hussey, P.J., Hutchison, C.J., 2007. A novel role for the nuclear membrane protein emerin in association of the

46

centrosome to the outer nuclear membrane. J. Cell Biol. 178, 897–904. https://doi.org/10.1083/jcb.200702026 Sampson, T.R., Weiss, D.S., 2014. Exploiting CRISPR/Cas systems for biotechnology. BioEssays News Rev. Mol. Cell. Dev. Biol. 36, 34–38. https://doi.org/10.1002/bies.201300135 Samson, C., Celli, F., Hendriks, K., Zinke, M., Essawy, N., Herrada, I., Arteni, A.-A., Theillet, F.- X., Alpha‐Bazin, B., Armengaud, J., Coirault, C., Lange, A., Zinn‐Justin, S., 2017. Emerin self-assembly mechanism: role of the LEM domain. FEBS J. 284, 338–352. https://doi.org/10.1111/febs.13983 Schirmer, E.C., Florens, L., Guan, T., Yates, J.R., Gerace, L., 2003. Nuclear membrane proteins with potential disease links found by subtractive proteomics. Science 301, 1380– 1382. https://doi.org/10.1126/science.1088176 Seto, E., Yoshida, M., 2014. Erasers of Histone Acetylation: The Histone Deacetylase Enzymes. Cold Spring Harb. Perspect. Biol. 6. https://doi.org/10.1101/cshperspect.a018713 Sivakumar, A., de las Heras, J.I., Schirmer, E.C., 2019. Spatial Genome Organization: From Development to Disease. Front. Cell Dev. Biol. 7. https://doi.org/10.3389/fcell.2019.00018 Sola Carvajal, A., McKenna, T., Wallén Arzt, E., Eriksson, M., 2015. Overexpression of Lamin B Receptor Results in Impaired Skin Differentiation. PLoS ONE 10. https://doi.org/10.1371/journal.pone.0128917 Solovei, I., Wang, A.S., Thanisch, K., Schmidt, C.S., Krebs, S., Zwerger, M., Cohen, T.V., Devys, D., Foisner, R., Peichl, L., Herrmann, H., Blum, H., Engelkamp, D., Stewart, C.L., Leonhardt, H., Joffe, B., 2013. LBR and lamin A/C sequentially tether peripheral heterochromatin and inversely regulate differentiation. Cell 152, 584–598. https://doi.org/10.1016/j.cell.2013.01.009 Sönnichsen, B., Koski, L.B., Walsh, A., Marschall, P., Neumann, B., Brehm, M., Alleaume, A.- M., Artelt, J., Bettencourt, P., Cassin, E., Hewitson, M., Holz, C., Khan, M., Lazik, S., Martin, C., Nitzsche, B., Ruer, M., Stamford, J., Winzi, M., Heinkel, R., Röder, M., Finell, J., Häntsch, H., Jones, S.J.M., Jones, M., Piano, F., Gunsalus, K.C., Oegema, K., Gönczy, P., Coulson, A., Hyman, A.A., Echeverri, C.J., 2005. Full-genome RNAi profiling of early embryogenesis in Caenorhabditis elegans. Nature 434, 462–469. https://doi.org/10.1038/nature03353 Steglich, B., Filion, G.J., van Steensel, B., Ekwall, K., 2012. The inner nuclear membrane proteins Man1 and Ima1 link to two different types of chromatin at the nuclear periphery in S. pombe. Nucl. Austin Tex 3, 77–87. https://doi.org/10.4161/nucl.18825 Stewart, C.L., Roux, K.J., Burke, B., 2007. Blurring the boundary: the nuclear envelope extends its reach. Science 318, 1408–1412. https://doi.org/10.1126/science.1142034 Stroud, H., Otero, S., Desvoyes, B., Ramírez-Parra, E., Jacobsen, S.E., Gutierrez, C., 2012. Genome-wide analysis of histone H3.1 and H3.3 variants in Arabidopsis thaliana. Proc. Natl. Acad. Sci. 109, 5370–5375. https://doi.org/10.1073/pnas.1203145109 Sullivan, T., Escalante-Alcalde, D., Bhatt, H., Anver, M., Bhat, N., Nagashima, K., Stewart, C.L., Burke, B., 1999. Loss of a-Type Lamin Expression Compromises Nuclear Envelope Integrity Leading to Muscular Dystrophy. J. Cell Biol. 147, 913–920. Szczerbal, I., Foster, H.A., Bridger, J.M., 2009. The spatial repositioning of adipogenesis genes is correlated with their expression status in a porcine mesenchymal stem cell

47

adipogenesis model system. Chromosoma 118, 647–663. https://doi.org/10.1007/s00412-009-0225-5 Szenker, E., Ray-Gallet, D., Almouzni, G., 2011. The double face of the histone variant H3.3. Cell Res. 21, 421–434. https://doi.org/10.1038/cr.2011.14 Taddei, A., Hediger, F., Neumann, F.R., Gasser, S.M., 2004. The Function of Nuclear Architecture: A Genetic Approach. Annu. Rev. Genet. 38, 305–345. https://doi.org/10.1146/annurev.genet.37.110801.142705 Takahashi, K., Yamanaka, S., 2006. Induction of pluripotent stem cells from mouse embryonic and adult cultures by defined factors. Cell 126, 663–676. https://doi.org/10.1016/j.cell.2006.07.024 Talwar, S., Kumar, A., Rao, M., Menon, G.I., Shivashankar, G.V., 2013. Correlated Spatio- Temporal Fluctuations in Chromatin Compaction States Characterize Stem Cells. Biophys. J. 104, 553–564. https://doi.org/10.1016/j.bpj.2012.12.033 Tamura, T., Smith, M., Kanno, T., Dasenbrock, H., Nishiyama, A., Ozato, K., 2009. Inducible deposition of the histone variant H3.3 in interferon-stimulated genes. J. Biol. Chem. 284, 12217–12225. https://doi.org/10.1074/jbc.M805651200 Thanisch, K., Song, C., Engelkamp, D., Koch, J., Wang, A., Hallberg, E., Foisner, R., Leonhardt, H., Stewart, C.L., Joffe, B., Solovei, I., 2017. Nuclear envelope localization of LEMD2 is developmentally dynamic and lamin A/C dependent yet insufficient for heterochromatin tethering. Differ. Res. Biol. Divers. 94, 58–70. https://doi.org/10.1016/j.diff.2016.12.002 Trojer, P., Reinberg, D., 2007. Facultative heterochromatin: is there a distinctive molecular signature? Mol. Cell 28, 1–13. https://doi.org/10.1016/j.molcel.2007.09.011 Udugama, M., M. Chang, F.T., Chan, F.L., Tang, M.C., Pickett, H.A., R. McGhie, J.D., Mayne, L., Collas, P., Mann, J.R., Wong, L.H., 2015. Histone variant H3.3 provides the heterochromatic H3 lysine 9 tri-methylation mark at telomeres. Nucleic Acids Res. 43, 10227–10237. https://doi.org/10.1093/nar/gkv847 Van de Vosse, D.W., Wan, Y., Wozniak, R.W., Aitchison, J.D., 2011. Role of the nuclear envelope on genome organization and gene expression. Wiley Interdiscip. Rev. Syst. Biol. Med. 3, 147–166. https://doi.org/10.1002/wsbm.101 van Steensel, B., Belmont, A.S., 2017. Lamina-associated domains: links with chromosome architecture, heterochromatin and gene repression. Cell 169, 780–791. https://doi.org/10.1016/j.cell.2017.04.022 Vijayaraghavan, B., Figueroa, R.A., Bergqvist, C., Gupta, A.J., Sousa, P., Hallberg, E., 2018. RanGTPase regulates the interaction between the inner nuclear membrane proteins, Samp1 and Emerin. Biochim. Biophys. Acta BBA - Biomembr. 1860, 1326–1334. https://doi.org/10.1016/j.bbamem.2018.03.001 Vijayaraghavan, B., Jafferali, M.H., Figueroa, R.A., Hallberg, E., 2016. Samp1, a RanGTP binding transmembrane protein in the inner nuclear membrane. Nucl. Austin Tex 7, 415–423. https://doi.org/10.1080/19491034.2016.1220465 Vlcek, S., Foisner, R., 2007. A-type lamin networks in light of laminopathic diseases. Biochim. Biophys. Acta BBA - Mol. Cell Res., Integrated approaches to cytoskeleton research 1773, 661–674. https://doi.org/10.1016/j.bbamcr.2006.07.002 Walther, T.C., Askjaer, P., Gentzel, M., Habermann, A., Griffiths, G., Wilm, M., Mattaj, I.W., Hetzer, M., 2003. RanGTP mediates nuclear pore complex assembly. Nature 424, 689–694. https://doi.org/10.1038/nature01898

48

Wang, Y., Lichter-Konecki, U., Anyane-Yeboa, K., Shaw, J.E., Lu, J.T., Östlund, C., Shin, J.-Y., Clark, L.N., Gundersen, G.G., Nagy, P.L., Worman, H.J., 2016. A mutation abolishing the ZMPSTE24 cleavage site in prelamin A causes a progeroid disorder. J. Cell Sci. 129, 1975–1980. https://doi.org/10.1242/jcs.187302 Wilkie, G.S., Korfali, N., Swanson, S.K., Malik, P., Srsen, V., Batrakou, D.G., de las Heras, J., Zuleger, N., Kerr, A.R.W., Florens, L., Schirmer, E.C., 2011. Several novel nuclear envelope transmembrane proteins identified in skeletal muscle have cytoskeletal associations. Mol. Cell. Proteomics MCP 10, M110.003129. https://doi.org/10.1074/mcp.M110.003129 Williams, R.R.E., Azuara, V., Perry, P., Sauer, S., Dvorkina, M., Jørgensen, H., Roix, J., McQueen, P., Misteli, T., Merkenschlager, M., Fisher, A.G., 2006. Neural induction promotes large-scale chromatin reorganisation of the Mash1 locus. J. Cell Sci. 119, 132–140. https://doi.org/10.1242/jcs.02727 Wilson, K.L., Berk, J.M., 2010. The nuclear envelope at a glance. J. Cell Sci. 123, 1973–1978. https://doi.org/10.1242/jcs.019042 Wolffe, A.P., 2001. Nucleosomes: Detailed Structure and Mutations, in: ELS. American Cancer Society. https://doi.org/10.1038/npg.els.0001156 Worman, H.J., Courvalin, J.-C., 2005. Nuclear envelope, nuclear lamina, and inherited disease. Int. Rev. Cytol. 246, 231–279. https://doi.org/10.1016/S0074- 7696(05)46006-4 Worman, H.J., Schirmer, E.C., 2015. Nuclear membrane diversity: underlying tissue-specific pathologies in disease? Curr. Opin. Cell Biol. 34, 101–112. https://doi.org/10.1016/j.ceb.2015.06.003 Yao, J., Fetter, R.D., Hu, P., Betzig, E., Tjian, R., 2011. Subnuclear segregation of genes and core promoter factors in myogenesis. Genes Dev. 25, 569–580. https://doi.org/10.1101/gad.2021411 Yates, J.R., Bagshaw, J., Aksmanovic, V.M., Coomber, E., McMahon, R., Whittaker, J.L., Morrison, P.J., Kendrick-Jones, J., Ellis, J.A., 1999. Genotype-phenotype analysis in X- linked Emery-Dreifuss muscular dystrophy and identification of a missense mutation associated with a milder phenotype. Neuromuscul. Disord. NMD 9, 159–165. Zhang, C., Clarke, P.R., 2001. Roles of Ran–GTP and Ran–GDP in precursor vesicle recruitment and fusion during nuclear envelope assembly in a human cell-free system. Curr. Biol. 11, 208–212. https://doi.org/10.1016/S0960-9822(01)00053-7 Zhang, J., Lian, Q., Zhu, G., Zhou, F., Sui, L., Tan, C., Mutalif, R.A., Navasankari, R., Zhang, Y., Tse, H.-F., Stewart, C.L., Colman, A., 2011. A human iPSC model of Hutchinson Gilford Progeria reveals vascular smooth muscle and mesenchymal stem cell defects. Cell Stem Cell 8, 31–45. https://doi.org/10.1016/j.stem.2010.12.002 Zheng, R., Ghirlando, R., Lee, M.S., Mizuuchi, K., Krause, M., Craigie, R., 2000. Barrier-to- autointegration factor (BAF) bridges DNA in a discrete, higher-order complex. Proc. Natl. Acad. Sci. U. S. A. 97, 8997–9002. https://doi.org/10.1073/pnas.150240197 Zuleger, N., Boyle, S., Kelly, D.A., de las Heras, J.I., Lazou, V., Korfali, N., Batrakou, D.G., Randles, K.N., Morris, G.E., Harrison, D.J., Bickmore, W.A., Schirmer, E.C., 2013. Specific nuclear envelope transmembrane proteins can promote the location of chromosomes to and from the nuclear periphery. Genome Biol. 14, R14. https://doi.org/10.1186/gb-2013-14-2-r14

49

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., Singh, H., 2012. DNA sequence-dependent compartmentalization and silencing of chromatin at the nuclear lamina. Cell 149, 1474–1487. https://doi.org/10.1016/j.cell.2012.04.035

50