Homozygous and Compound Heterozygous Mutations in ZMPSTE24 Cause the Laminopathy Restrictive Dermopathy See related Commentary on page xii Homozygous and Compound Heterozygous Mutations in ZMPSTE24 Cause the Laminopathy Restrictive Dermopathy Casey L. Moulson,� Gloriosa Go,� Jennifer M. Gardner,w Allard C. van der Wal,z J. Henk Sillevis Smitt,y Johanna M. van Hagen,z and Jeffrey H. Miner�# �Department of Internal Medicine, Renal Division and wDepartment of Genetics, Washington University School of Medicine, St Louis, Missouri, USA; zDepartment of Cardiovascular Pathology and yDepartment of Dermatology, Academic Medical Center, Amsterdam, The Netherlands; zDepartment of Clinical Genetics and Human Genetics, VU University Medical Center, Amsterdam, The Netherlands; #Department of Cell Biology and Physiology, Washington University School of Medicine, St Louis, Missouri, USA Restrictive dermopathy (RD) is a lethal human genetic disorder characterized by very tight, thin, easily eroded skin, rocker bottom feet, and joint contractures. This disease was recently reported to be associated with a single heterozygous mutation in ZMPSTE24 and hypothesized to be a digenic disorder (Navarro et al, Lamin A and ZMPSTE24 (FACE-1) defects cause nuclear disorganization and identify restrictive dermopathy as a lethal neonatal laminopathy. Hum Mol Genet 13:2493–2503, 2004). ZMPSTE24 encodes an enzyme necessary for the correct processing and maturation of lamin A, an intermediate filament component of the nuclear envelope. Here we present four unrelated patients with homozygous mutations in ZMPSTE24 and a fifth patient with compound heterozygous mutations in ZMPSTE24. Two of the three different mutations we found are novel, and all are single base insertions that result in messenger RNA frameshifts. As a consequence of the presumed lack of ZMPSTE24 activity, prelamin A, the unprocessed toxic form of lamin A, was detected in the nuclei of both cultured cells and tissue from RD patients, but not in control nuclei. Abnormally aggregated lamin A/C was also observed. These results indicate that RD is an autosomal recessive laminopathy caused by inactivating ZMPSTE24 mutations that result in defective processing and nuclear accumulation of prelamin A. Key words: FATP4 protein/lamin A/nuclear envelope/STE24 protein J Invest Dermatol 125:913 – 919, 2005 Restrictive dermopathy (RD), also known as lethal tight skin contracture syndrome (MIM 275210), is a rare genetic dis- order that results in death, usually within several hours or days of birth. Babies with RD have tight, translucent, par- tially eroded skin, joint contractures, rocker bottom feet, and distinctive craniofacial abnormalities that typically in- clude micrognathia, a facial expression with the mouth fixed in an ‘‘O’’ position, a small pinched nose, and low-set ears (Welsh et al, 1992; Mau et al, 1997; Sillevis Smitt et al, 1998). In most patients, death results from respiratory distress as a result of severely restricted movements. Of the few RD cases reported in the literature, many arose in children of consanguineous parents, and the disease re- curs within families; its mode of inheritance had therefore been presumed to be autosomal recessive (Welsh et al, 1992; Sillevis Smitt et al, 1998). Recently, however, Navarro et al (2004) used a candidate gene approach to identify a single inactivating heterozygous mutation in ZMPSTE24 (c.1085dupT) associated with RD in six unrelated, noncon- sanguineous families of French and Dutch heritage. ZMPSTE24 encodes a zinc metalloproteinase required for lamin A maturation (Bergo et al, 2002; Pendas et al, 2002), suggesting that RD is a laminopathy that affects the structure of the nucleus. Because only one allele was found to be mutated in each affected individual, and the parents carrying that allele were normal, the authors hypothesized a digenic mode of inheritance for RD (Navarro et al, 2004). Contrary to this, we have identified four patients with three different homozygous mutations and one patient with com- pound heterozygous mutations in ZMPSTE24 that are as- sociated with RD, providing strong support for a simple autosomal recessive mode of inheritance. We also show that in two patients with different homozygous mutations, there are nuclear abnormalities associated with accumula- tion of unprocessed prelamin A, thereby providing novel evidence that RD is a laminopathy. Results Identification of ZMPSTE24 mutations associated with RD We became interested in RD upon discovering a spontaneous, autosomal recessive mouse mutation that we named wrinkle free (wrfr). wrfr �/� mice have very tight, thick skin, and their mouths are fixed in an open position (Moulson et al, 2003). They die within 24 h of birth from an inability to breathe properly or to suckle, because of Abbreviations: DHPLC, denaturing high-performance liquid chro- matography; FACE, farnesyiated protein-converting enzyme 1; mRNA, messenger RNA; PBS, phosphate-buffered saline; PCR, polymerase chain reaction; RD, restrictive dermopathy; SNP, single nucleotide polymorphism Copyright r 2005 by The Society for Investigative Dermatology, Inc. 913 severely restricted movements. The overall phenotype is thus reminiscent of RD. Positional cloning of wrfr revealed a retrotransposon insertion in Slc27a4, which encodes fatty acid transport protein 4 (Moulson et al, 2003). In an effort to determine whether SLC27A4 plays a role in human RD, we obtained blood and/or tissues from mem- bers of six RD kindreds, three of which were known or suspected to involve consanguinity. DNA was prepared from individuals with RD, all of whom died within hours or days of birth and exhibited the classic features described by Witt (Witt et al, 1986). DNA was also prepared from parents and any unaffected full siblings when available. We se- quenced SLC27A4 exons and intron–exon junctions in three unrelated patients but found no mutations. To formally rule out SLC27A4 as a candidate, and to identify potential can- didate regions, we performed genome-wide single nucleo- tide polymorphism (SNP) genotyping using Affymetrix GeneChip Mapping arrays (Matsuzaki et al, 2004) on a small but potentially informative subset of these DNA (data not shown). Using DNA-Chip (dCHIP) software (Lin et al, 2004) to display the results, and assuming an autosomal recessive mode of inheritance, we identified a 15 Mb locus on chromosome 1p32–35 as a major RD candidate region, whereas the SLC27A4 locus at 9q34 was ruled out. Con- sistent with the report of Navarro et al, 1p32–35 contains ZMPSTE24. Therefore, we sequenced ZMPSTE24 exons and flanking intronic sequences amplified by polymerase chain reaction (PCR) from RD family members’ DNA to search for mutations and to determine their association with RD. A consanguineous Dutch family had one child affected with RD and five unaffected children. We identified in this family the same c.1085dupT insertion reported by Navarro et al (2004) (Fig 1A and B). In contrast to the findings of Navarro et al, however, here the mutation was homozygous in the affected patient, whereas both parents were het- erozygous (Fig 1A and B). DNA was also available from four of the five unaffected siblings; three were heterozygous for the insertion, and one did not carry it. We found this same mutation in a nonconsanguineous American family of Ger- man ancestry (Fig 1C). The mutation was homozygous in one affected child, and both parents and one brother were heterozygous; the other brother did not carry the mutation. No DNA from the second affected child was available. In this family, the apparently identical mutations were linked to different genotypes at a novel SNP in ZMPSTE24 intron 4 (data not shown), suggesting that the mutations arose independently. Analysis of ZMPSTE24 in a consanguineous Guatemalan family with two affected children revealed a novel single base duplication in exon 5 (c.591dupT) (Fig 1D and E). This mutation was homozygous in one affected child and het- erozygous in both parents (Fig 1D and E). DNA from the second child was not available. A novel single base duplication was present in ZMPSTE24 exon 1 (c.54dupT) in a Mennonite kindred with five affected children out of six born to two families (Fig 1F and G). DNA was only available from one affected child; the mutation was homozygous in her and heterozygous in both her parents. The other parents and their unaffected son were heterozygous for the same mutation. RD has been reported in a Mennonite kindred before (Lowry et al, 1985), suggesting that this unique mutation is segregating in the Mennonite population. We also assayed DNA extracted from archived paraffin- embedded tissues taken from a Dutch RD patient. Two dif- ferent heterozygous mutations were detected: the exon 5 mutation found in the Guatemalan family (c.591dupT) and the previously identified exon 9 insertion (c.1085dupT). We conclude that this patient is compound heterozygous, al- though we cannot formally prove it because parental DNA are not available. Might these mutations actually be common polymorph- isms? The c.1085dupT mutation in exon 9 has been pre- viously reported, and it was not found to be present on 300 control chromosomes (Agarwal et al, 2003; Navarro et al, 2004). Denaturing high-performance liquid chromatography (DHPLC) analysis of exons 1 and 5 amplified by PCR from 50 CEPH grandparent DNA showed no evidence of the c.54dupT and c.591dupT mutations, respectively (data not shown). Searches of the Celera Discovery System (http://myscience.appliedbiosystems.com/) and public SNP (http://www.ncbi.nlm.nih.gov/SNP/) databases did not re- veal any insertions in the ZMPSTE24 coding region, al- though 470 SNP have been identified in the gene. Together with the fact that all three sequence variants shift the mes- senger RNA (mRNA) reading frame, these findings suggest that they are genuine pathogenic mutations. Consequences of the mutations on expression of ZMPSTE24 All three mutations are single base duplications in ZMPSTE24 coding exons. They are therefore frameshift mutations that would be expected to lead to production of truncated versions of ZMPSTE24, which normally consists of 475 amino acids. c.54dupT would encode p.Ile19- TyrfsX28; c.591dupT would encode p.Ile198TyrfsX20; and c.1085dupT would encode p.Leu362PhefsX19, according to standard nomenclature (den Dunnen and Antonarakis, 2000; see also http://www.genomic.unimelb.edu.au/mdi/ mutnomen/recs.html). These truncated proteins would be expected to have little if any normal function, as ZMPSTE24 is conserved in terms of both sequence and enzymatic activity from yeast to humans. Indeed, human ZMPSTE24 can complement the yeast ste24 mutation (Schmidt et al, 2000), but p.Leu362PhefsX19, which is en- coded by the c.1085dupT mutation, is noncomplementing (Agarwal et al, 2003). Aside from encoding truncated proteins, the mutant transcripts might also be subject to nonsense-mediated mRNA decay (NMD) (Hentze and Kulozik, 1999), which could preclude synthesis of any truncated ZMPSTE24 pro- teins. To assay for NMD, we analyzed RNA isolated from fibroblasts derived from the American patient carrying the c.1085dupT mutation and from a control. Northern blotting did not reveal a significant decrease in ZMPSTE24 mRNA levels in the RD fibroblasts (data not shown), suggesting that NMD was not a major factor. One possibility to explain this is that exon 9 containing the frameshift might be skipped, leading to an in-frame splice from exon 8 to exon 10 and elimination of the premature stop codon. But reverse transcriptase-polymerase chain reaction (RT-PCR) analysis did not reveal any evidence for skipping of exon 9 (Fig 2A). 914 MOULSON ET AL THE JOURNAL OF INVESTIGATIVE DERMATOLOGY RNA was also prepared from the liver of the Guatemalan patient, but it was highly degraded and not suitable for northern blotting. Nevertheless, RT-PCR analysis revealed that there was no in-frame skipping of the mutated exon 5 (Fig 2A), which could have led to production of a protein. Consistent with these RT-PCR data, western blotting with a ZMPSTE24 COOH-terminal antibody detected little if any ZMPSTE24 in patient fibroblasts or liver, but an appropriate �50 kDa band was observed in controls (Fig 2B). Consequences of the mutations on localization of nu- clear envelope proteins The only known substrate for ZMPSTE24 is lamin A, encoded by the LMNA gene. This gene also encodes lamin C, a shorter form that is translated Figure 1 Family pedigrees, ZMPSTE24 genotypes, and sequence chromatograms. Arrows indicate individuals whose DNA was tested. ZMPSTE24 mutations are shown below the pedigree symbols; all are hypothesized to cause loss of activity. (A) Pedigree of a consanguineous Dutch family; the father’s maternal grandmother and the mother’s paternal grandmother were sisters. The affected child had a homozygous thymidine duplication in ZMPSTE24 exon 9. (B) Sequence chromatograms of exon 9 (reverse direction) show a duplicated thymidine in a stretch of nine. Overlap of the normal and mutated strand sequences occurred after the insertion in this and in all the other heterozygous samples described. (C) Pedigree of a nonconsanguineous American family carrying the same mutation as the Dutch family. (D) Pedigree from a consanguineous Guatemalan family; the father’s paternal grandfather was a paternal uncle of the mother. (E) Sequence chromatograms of exon 5 (forward direction) show a duplicated thymidine in a stretch of seven. (F) Five affected children were born to two pairs of related parents from a Mennonite kindred; the fathers were brothers of each other, and the mothers were first cousins of each other. (G) Sequence chromatograms of ZMPSTE24 exon 1 (reverse direction) show an extra thymidine homozygous in the affected child. ZMPSTE24 MUTATIONS IN RESTRICTIVE DERMOPATHY 915125 : 5 NOVEMBER 2005 from an alternatively spliced mRNA. Lamin A maturation requires ZMPSTE24 activity, but lamin C maturation does not. Lamin A/C distribution is abnormal in RD patients re- ported to have a heterozygous c.1085dupT mutation in ZMPSTE24 (Navarro et al, 2004). Because the c.1085dupT mutation in the patient from whom we obtained fibroblasts was homozygous, we examined whether there was aberrant distribution of lamin A/C using an antibody that recognizes both (Fig 3A, B). Lamin A/C appeared to be distributed in clusters in the nucleus, as opposed to the homogeneous, smooth, uniform distribution in nuclei from normal fibroblast cultures. Accumulation of unprocessed prelamin A, which is toxic (Fong et al, 2004), has been reported in Zmpste24 knockout mice (Bergo et al, 2002; Pendas et al, 2002). To examine whether there was accumulation of prelamin A in RD pa- tients as well, cultured fibroblasts were stained with a pre- lamin A-specific antibody. Strong prelamin A staining was observed in RD fibroblast nuclei (Fig 3D), whereas no pre- lamin A was detected in control fibroblasts (Fig 3C). Nuclear prelamin A was also observed in frozen liver sections from the Guatemalan patient (Fig 3F) but not in liver from a con- trol, although lamin A/C was clearly present (Fig 3E). The presence of prelamin A in patient but not control liver was confirmed by western blotting (Fig 2C). We conclude that RD is associated with a specific defect in prelamin A processing caused by the absence of functional ZMPSTE24. Discussion In this study, we identified two novel mutations and a third previously described mutation in ZMPSTE24 that are as- sociated with RD. Importantly, we also show that RD is a simple autosomal recessive disorder. Though ZMPSTE24 had already been implicated in RD, only one heterozygous mutation was identified, and it was also present in unaf- fected family members, leading to the conclusion that RD Figure 2 Analysis of ZMPSTE24 messenger RNA (mRNA) and protein in control and restrictive dermopathy (RD) samples. (A) Reverse tran- scriptase polymerase chain reaction analysis of total RNA from liver and fibroblasts reveals no evidence of splicing around the exons con- taining the c.591dupT (exon 5; lanes 1 and 2) and c.1085dupT (exon 9; lanes 5 and 6) mutations. Glyceraldehyde-3-phosphate dehydrogenase primers were used as a control in lanes 3, 4, 7, and 8. Control fetal liver, lanes 1 and 3; c.591dupT liver, lanes 2 and 4; control fibroblasts, lanes 5 and 7; c.1085dupT fibroblasts, lanes 6 and 8. (B) Western blot analysis of ZMPSTE24 (top) and b-actin (bottom). ZMPSTE24 was detected in control but not RD extracts; the same blot reprobed for b-actin shows protein in all lanes. Lane 1, control fetal liver; Lane 2, RD liver (Gua- temalan patient); Lane 3, control fibroblasts; Lane 4, RD fibroblasts (American patient). (C) Western blot analysis of prelamin A and b-actin. Prelamin A ( � 70 kDa) was detected in liver from the RD patient (lane 2) but not from the control (lane 1). The blot was reprobed for b-actin to show that protein was present in both lanes. Figure 3 Immunofluorescent localization of lamin A/C and prelamin A in control and restrictive dermopathy (RD) samples. The organization of lamins in fibroblast nuclear envelopes is disrupted by accumulation of prelamin A in the absence of ZMPSTE24. Confocal micrographs show localization of lamins A and C in nuclei of a control fibroblast (A) and a fibroblast from the American RD patient (B). Prelamin A was undetectable in control fibroblasts (C), but there was nuclear accumu- lation in RD patient fibroblasts (D). In liver sections, prelamin A was undetectable in the control (E) but clearly present in RD patient nuclei (F). The control section did contain nuclei, which are stained for lamin A/C (inset in E). 916 MOULSON ET AL THE JOURNAL OF INVESTIGATIVE DERMATOLOGY was a digenic disorder (Navarro et al, 2004). Rather than there being a digenic mode of inheritance, we favor the pos- sibility that there exist ZMPSTE24 mutations that are diffi- cult to detect by conventional PCR-based methods. This is perhaps the most straightforward explanation for how there could be an absence of mature lamin A even with one ap- parently wild-type copy of ZMPSTE24. Such mutations could lie in the gene’s promoter, or could involve segmental deletions that are complemented—only as far as PCR is concerned—by normal regions of the other allele. Indeed, we could not find any mutations in a pair of affected fra- ternal twins from Finland, though SNP genotyping showed that the twins share genotypes on the segment of chromo- some 1 containing the ZMPSTE24 locus (data not shown). All mutations identified thus far to cause RD are single base duplications that result in mRNA frameshifts, which likely prevent production of a functional protein. One mis- sense ZMPSTE24 mutation (c.1018T4C; p.Trp340Arg) has been found, however, in a Belgian patient with severe man- dibuloacral dysplasia with type B lipodystrophy (MADB— MIM 608612) (Agarwal et al, 2003). This patient was actually compound heterozygous; the other allele bore the same c.1085dupT-inactivating mutation that Navarro et al (2004) and we have subsequently found to be associated with RD. The mutant protein carrying the p.Trp340Arg substitution is partially active in yeast complementation assays, whereas p.Leu362PhefsX19, which is encoded by c.1085dupT, is totally inactive (Agarwal et al, 2003). Thus, the limited infor- mation available regarding ZMPSTE24 mutations suggests that inactivating mutations cause RD, whereas partial loss of function mutations are associated with less severe dis- ease, as is MADB. Identification of mutations in additional patients will help determine whether a definitive genotype/ phenotype correlation exists. Accumulation of prelamin A has been reported in Zmpste24 knockout mice (Bergo et al, 2002; Pendas et al, 2002) and in cultured cells from human patients with Hut- chinson–Gilford progeria syndrome (HGPS) (MIM 176670) (Goldman et al, 2004); its presence has also been hypoth- esized but not proved in other human laminopathies (Agarwal et al, 2003; Navarro et al, 2004). Our study dem- onstrates accumulation of prelamin A in affected human tissue in situ (Fig 3F), suggesting that it is not merely a consequence of cell culture. Accumulation of prelamin A has recently been shown to be toxic; disease phenotypes in Zmpste24 �/� mice are ‘‘rescued’’ by only a 50% reduction in the amount of prelamin A, via knockout of one Lmna allele (Fong et al, 2004). The presence of large amounts of pre- lamin A may also account for the increased severity of RD as compared with laminopathies with either lamin A muta- tions or a missense mutation in ZMPSTE24. Given the lack of detectable nuclear prelamin A in normal cells, it is pos- sible that a simple immunofluorescence assay for prelamin A in amniocyte and/or chorionic villi cell nuclei, in conjunc- tion with DNA analysis, might serve as a definitive prenatal test for RD and perhaps for other laminopathies. Preim- plantation genetic diagnosis (Sermon et al, 2004) should also be possible. Further studies will be necessary to de- termine the feasibility of these approaches. We found lamins A and C clustered and distributed in aggregates in the nuclei of RD patient fibroblasts, whereas normal fibroblast nuclei showed a homogeneous and smooth, even distribution, concentrated at the nuclear periphery. Abnormal distribution of lamins A and C in foci, aggregations, and honeycombs has been reported in a number of human patients carrying lamin A/C gene mutations (Novelli et al, 2002; Capanni et al, 2003; Caux et al, 2003; Muchir et al, 2004) as well as in cells transfected with lamin A/C mutants (Ostlund et al, 2001; Holt et al, 2003). The mechanism whereby mutations in LMNA and ZMPSTE24 that result in aberrant lamin A processing cause disease has not been determined. Several lines of evidence suggest that it is likely to be a combination of impaired nuclear stability and the resulting nuclear deformations (Lammerding et al, 2004) causing secondary alterations in heterochromatin localization and significant changes in gene expression (Nikolova et al, 2004). (For a review of these hypotheses, see reference Worman and Courvalin (2004).) In addition, lamin A binds directly or indirectly to a number of nuclear proteins, including emerin (Lee et al, 2001; Sakaki et al, 2001) and nesprin (Mislow et al, 2002), and attach them to the nuclear envelope. Lamin A also binds to transcription factors such as MOK2 (Dreuillet et al, 2002), retinoblastoma (Mancini et al, 1994), and SREBP1 (Lloyd et al, 2002). (See (Zastrow et al, 2004) for a review of lamin-binding proteins.) Mislocalization of lamin A, as we and others have observed, is likely indicative that some of its binding partners are also mislocalized and not bound to the nuclear envelope, which may have profound secondary effects on cells. Interestingly, the neonatally lethal wrinkle-free phenotype of Slc27a4 �/� mice that we and others have reported (Herrmann et al, 2003; Moulson et al, 2003) appears to be much more similar to the human RD phenotype than that exhibited by Zmpste24 �/� mice, which survive for several months. This may be coincidence, but the possibility exists that there is some special mechanistic relationship in humans between nuclear architecture and expression of genes involved in fatty acid homeostasis. Indeed, one of the features of the diverse laminopathies that are less severe than RD is lipodystrophy. Genetic and allelic heterogeneity appear to explain in part the phenotypic differences among the related lamino- pathies HGPS, MAD, and most recently, RD. Navarro et al included in their study two patients several months old with a skin disease less severe than typical RD, and they found one novel and one previously reported mutation in the LMNA gene. These are both likely to be sporadic, het- erozygous, dominant-negative mutations. The latter muta- tion had been found previously in several patients with HGPS. There is clearly a marked difference in disease se- verity between ZMPSTE24-based neonatally lethal RD and LMNA-based progeria. To avoid confusion, we propose that the descriptor ‘‘restrictive dermopathy’’ be reserved for individuals with the severe, neonatally lethal disease de- scribed by Witt (Witt et al, 1986). Children with some RD-like features who survive well past the neonatal period are more likely to manifest features of progeria or MAD and should be described as such. The ability to screen for ZMPSTE24 and LMNA mutations may make these distinctions possible at the molecular level. ZMPSTE24 MUTATIONS IN RESTRICTIVE DERMOPATHY 917125 : 5 NOVEMBER 2005 Materials and Methods Amplification of genomic DNA, DNA sequencing, and DHPLC The Washington University School of Medicine Human Studies Committee approved this study, which was conducted according to the Declaration of Helsinki principles. Informed pa- rental consent was obtained in all cases, except for one that in- volved archived tissues; for this a waiver of consent was approved. Genomic DNA was extracted from the blood of six affected indi- viduals from five different kindreds and from their parents and available unaffected siblings using QiaAMP DNA Blood Kit (Qiagen Inc., Valencia, California). DNA was extracted from paraffin-em- bedded tissues from a seventh affected individual as described (Coombs et al, 1999). Exons and intron–exon junctions from genomic DNA samples were amplified using KlentaqLA (BD Bio- sciences Clontech, Palo Alto, California). Each 20 mL reaction contained 1 � KLA (KlentaqLA) buffer, 125 mM dNTP (deoxribo- nucleotide triphosphates) 2.5 mM MgCl2, 10 pmoles of each prim- er, 20 ng genomic DNA, and 1 U KlentaqLA. Cycling conditions were a 3 min initial denaturation at 951C, followed by 35 cycles of 30 s at 941C, 1 min at 571C, and 3 min at 701C. PCR products were gel purified and recovered using the Wizard SV Gel Clean-up Sys- tem (Promega, Madison, Wisconsin). The Washington University Protein and Nucleic Acid Chemistry Lab performed automated sequencing reactions using internal primers with Big-Dye Termi- nator 3.1 and an Applied Biosystems 3730 DNA sequencer (Ap- plied Biosystems, Foster City, California). The PCR and sequencing primer sequences used are shown in Table SI. For DHPLC, exons 1 and 5 were amplified from 50 CEPH individuals and from heterozygous parents as described above. Following denaturation and slow reannealing, amplicons were subjected to DHPLC analysis on a Transgenomic WAVE System containing a DNASepHT column (Transgenomic, Omaha, Nebras- ka) in a linear acetonitrile gradient. RNA purification and RT-PCR Total RNA was isolated from liver homogenized in Tri-Reagent (Molecular Research Center, Cincin- nati, Ohio) and from fibroblasts scraped into Tri-Reagent according to the manufacturer’s instructions. 0.5 mg of total RNA was reverse transcribed using Superscript III (Invitrogen, Carlsbad, California) with oligo(dT)12–18 primers in a 20 mL reaction. Touchdown PCR was performed in a 20 mL reaction using 1 mL of complementary DNA. PCR reactions contained manufacturer-supplied 1 � reac- tion buffer, 2.5 mM magnesium chloride, 0.125 mM of each dNTP, 10 pmoles of each primer, and 0.5 U Biolase DNA polymerase (MidSci, St Louis, Missouri). Primers are shown in Table SII. Western blotting Proteins were extracted from liver by sonication in 8 M urea, 10 mM tris pH 8.0, 1 mM ethylenediaminetetraacetic acid, and Complete Protease Inhibitors (Roche Diagnostics, India- napolis, Indiana). Fibroblasts were scraped into cold phosphate- buffered saline (PBS) containing protease inhibitors, pelleted, resuspended in the urea buffer, and sonicated. Proteins were separated on 8% sodium dodecyl sulfate-polyacrylamide gels and blotted on to nitrocellulose. Blots were incubated in 5% nonfat dry milk in TBST (Tris-buffered saline/0.02% Tween-20) for 1 h prior to incubation with primary antibodies overnight at 41C. Primary antibodies were goat anti-human prelamin A (sc-6214; Santa Cruz Biotechnology, Santa Cruz, California), rabbit anti-human ZMPSTE24 (AP2415b; Abgent, San Diego, California), and rabbit anti-human b-actin (4967; Cell Signaling Technology, Beverly, Mas- sachusetts). The secondary antibody for b-actin was conjugated to horseradish peroxidase, whereas secondary antibodies for pre- lamin A and ZMPSTE24 were detected with avidin–biotin and horseradish peroxidase. Visualization was by enhanced chemilu- minescence (Amersham, Arlington Heighst, Illinois). Cell culture and immunofluorescence Skin fibroblasts from the American RD patient and normal skin fibroblasts (Detroit 551 from the American Type Culture Collection, Manassas, Virginia) were cultured in Earle’s Minimal Essential medium modified to contain 2 mM L-glutamine, 1 mM sodium pyruvate, 0.1 mM nonessential amino acids, 1.5 g per liter sodium bicarbonate, and 10% fetal bovine serum. For immunostaining, cells were cultured for several days on glass Falcon BioCoat CultureSlides (BD Biosciences Dis- covery Labware, Bedford, Massachusetts). Cells were fixed for 10 min at 41 C in 2% fresh paraformaldehyde in PBS, washed in PBS, permeabilized for 10 min in cold methanol, and air-dried. Liver tissue from the Guatemalan RD patient and from a control were frozen in OCT, sectioned on a cryostat, air-dried for 30 min, and fixed and permeabilized as described for cultured cells. For immunofluorescence, samples were blocked in either 4% bovine serum albumin (BSA) in PBS or 1% heat-inactivated normal goat serum in PBS. Antibodies were as follows: anti-lamins A and C (MAB3538) and fluorescein isothiocyanate-conjugated anti-goat immunoglobulin G (IgG) from Chemicon (Temecula, California); goat anti-prelamin A; and Alexa 488-conjugated anti-mouse IgG1 from Molecular Probes (Eugene, Oregon). All antibody dilutions were in PBS containing 1% BSA, and washes were in PBS. Fol- lowing primary and secondary antibody incubations at room tem- perature for 1 h each and washing, slides were mounted in 1 mg per mL p-phenylenediamine/0.1 � PBS/90% glycerol. Confocal microscope images were obtained using a Zeiss compound microscope (Carl Zeiss International, Jena, Germany) with a BioRad MRC 1024 confocal adaptor (BioRad Laboratories, Inc., Hercules, California). All other micrographs were obtained with a Spot 2 cooled color digital camera (Diagnostic Instruments, Sterling Heights, Michigan) attached to a Nikon Eclipse E800 compound microscope (Nikon USA, Melville, New York). Images were imported into Adobe Photoshop 5 and Adobe Illustrator 9 for processing and layout. We are extremely grateful to the families who participated in this study and to the referring physicians, scientists, and genetic counselors for their invaluable contributions. We thank The Laboratory of Develop- mental Biology at the Univerisity of Washington in Seattle, supported by National Institutes of Health grant R24HD000836, for supplying human fetal liver tissue, and Edward Fox and the Dana-Farber Cancer Institute Microarray Core for performing SNP genotyping; the Wash- ington University School of Medicine Division of Human Genetics Genotyping Core for facilitating DHPLC analysis; Martin Pollak and Alan Beggs for advice; Phillip Stahl and Didier Hodzic for use of the confocal microscope; and Paul Goodfellow for helpful suggestions and comments on the manuscript. This work was funded by National In- stitutes of Health grant R01AR049269 to JHM. Supplementary Material The following material is available online for this article. Table S1 Primers used to amplify and sequence ZMPSTE24 exons and flanking itronic sequences. Table S2 Primers used for RT-PCR. DOI: 10.1111/j.0022-202X.2005.23846.x Manuscript received February 4, 2005; revised April 11, 2005; accept- ed for publication May 3, 2005 After this paper was accepted, we found a novel ZMPSTE24 nonsense mutation, c.691G4T, predicted to encode p.Glu231X, in both parents of an affected child from southern India. DNA from the child was not available, but we consider this to be a pathogenic mutation. Navarro et al. recently published a paper (Loss of ZMPSTE24 (FACE-1) causes autosomal recessive restrictive dermopathy and accumulation of Lamin A precursors. Hum Mol. Genet 14:1503–1513, 2005) showing that both alleles of ZMPSTE24 are in fact mutated in their patients with restrictive dermopathy; these patients were previously reported to carry only heterozygous mutations. Based on these new findings, they have withdrawn their hypothesis of digenic inheritance. Address correspondence to: Dr Jeffrey H. Miner, Renal Division, Box 8126, Washington University School of Medicine, 660 S. Euclid Ave., St Louis, Missouri 63110, USA. Email: minerj@wustl.edu 918 MOULSON ET AL THE JOURNAL OF INVESTIGATIVE DERMATOLOGY References Agarwal AK, Fryns JP, Auchus RJ, Garg A: Zinc metalloproteinase, ZMPSTE24, is mutated in mandibuloacral dysplasia. Hum Mol Genet 12:1995–2001, 2003 Bergo MO, Gavino B, Ross J, et al: Zmpste24 deficiency in mice causes spon- taneous bone fractures, muscle weakness, and a prelamin A processing defect. Proc Natl Acad Sci USA 99:13049–13054, 2002 Capanni C, Cenni V, Mattioli E, et al: Failure of lamin A/C to functionally assemble in R482L mutated familial partial lipodystrophy fibroblasts: Altered inter- molecular interaction with emerin and implications for gene transcription. Exp Cell Res 291:122–134, 2003 Caux F, Dubosclard E, Lascols O, et al: A new clinical condition linked to a novel mutation in lamins A and C with generalized lipoatrophy, insulin-resistant diabetes, disseminated leukomelanodermic papules, liver steatosis, and cardiomyopathy. J Clin Endocrinol Metab 88:1006–1013, 2003 Coombs NJ, Gough AC, Primrose JN: Optimisation of DNA and RNA extraction from archival formalin-fixed tissue. Nucleic Acids Res 27:e12, 1999 den Dunnen JT, Antonarakis SE: Mutation nomenclature extensions and sug- gestions to describe complex mutations: A discussion. Hum Mutat 15:7–12, 2000 Dreuillet C, Tillit J, Kress M, Ernoult-Lange M: In vivo and in vitro interaction between human transcription factor MOK2 and nuclear lamin A/C. Nu- cleic Acids Res 30:4634–4642, 2002 Fong LG, Ng JK, Meta M, et al: Heterozygosity for Lmna deficiency eliminates the progeria-like phenotypes in Zmpste24-deficient mice. Proc Natl Acad Sci USA 101:18111–18116, 2004 Goldman RD, Shumaker DK, Erdos MR, et al: Accumulation of mutant lamin A causes progressive changes in nuclear architecture in Hutchinson–Gilford progeria syndrome. Proc Natl Acad Sci USA 101:8963–8968, 2004 Hentze MW, Kulozik AE: A perfect message: RNA surveillance and nonsense- mediated decay. Cell 96:307–310, 1999 Herrmann T, van der Hoeven F, Grone HJ, et al: Mice with targeted disruption of the fatty acid transport protein 4 (Fatp 4, Slc27a4) gene show features of lethal restrictive dermopathy. J Cell Biol 161:1105–1115, 2003 Holt I, Ostlund C, Stewart CL, Man N, Worman HJ, Morris GE: Effect of path- ogenic mis-sense mutations in lamin A on its interaction with emerin in vivo. J Cell Sci 116:3027–3035, 2003 Lammerding J, Schulze PC, Takahashi T, et al: Lamin A/C deficiency causes defective nuclear mechanics and mechanotransduction. J Clin Invest 113: 370–378, 2004 Lee KK, Haraguchi T, Lee RS, Koujin T, Hiraoka Y, Wilson KL: Distinct functional domains in emerin bind lamin A and DNA-bridging protein BAF. J Cell Sci 114:4567–4573, 2001 Lin M, Wei LJ, Sellers WR, Lieberfarb M, Wong WH, Li C: dChipSNP: Significance curve and clustering of SNP-array-based loss-of-heterozygosity data. Bioinformatics 20:1233–1240, 2004 Lloyd DJ, Trembath RC, Shackleton S: A novel interaction between lamin A and SREBP1: Implications for partial lipodystrophy and other laminopathies. Hum Mol Genet 11:769–777, 2002 Lowry RB, Machin GA, Morgan K, Mayock D, Marx L: Congenital contractures, edema, hyperkeratosis, and intrauterine growth retardation: A fatal syndrome in Hutterite and Mennonite kindreds. Am J Med Genet 22: 531–543, 1985 Mancini MA, Shan B, Nickerson JA, Penman S, Lee WH: The retinoblastoma gene product is a cell cycle-dependent, nuclear matrix-associated pro- tein. Proc Natl Acad Sci USA 91:418–422, 1994 Matsuzaki H, Loi H, Dong S, et al: Parallel genotyping of over 10,000 SNPs using a one-primer assay on a high-density oligonucleotide array. Genome Res 14:414–425, 2004 Mau U, Kendziorra H, Kaiser P, Enders H: Restrictive dermopathy: Report and review. Am J Med Genet 71:179–185, 1997 Mislow JM, Holaska JM, Kim MS, Lee KK, Segura-Totten M, Wilson KL, McNally EM: Nesprin-1alpha self-associates and binds directly to emerin and lamin A in vitro. FEBS Lett 525:135–140, 2002 Moulson CL, Martin DR, Lugus JJ, Schaffer JE, Lind AC, Miner JH: Cloning of wrinkle-free, a previously uncharacterized mouse mutation, reveals cru- cial roles for fatty acid transport protein 4 in skin and hair development. Proc Natl Acad Sci USA 100:5274–5279, 2003 Muchir A, Medioni J, Laluc M, et al: Nuclear envelope alterations in fibroblasts from patients with muscular dystrophy, cardiomyopathy, and partial lipo- dystrophy carrying lamin A/C gene mutations. Muscle Nerve 30:444–450, 2004 Navarro CL, De Sandre-Giovannoli A, Bernard R, et al: Lamin A and ZMPSTE24 (FACE-1) defects cause nuclear disorganization and identify restrictive dermopathy as a lethal neonatal laminopathy. Hum Mol Genet 13:2493– 2503, 2004 Nikolova V, Leimena C, McMahon AC, et al: Defects in nuclear structure and function promote dilated cardiomyopathy in lamin A/C-deficient mice. J Clin Invest 113:357–369, 2004 Novelli G, Muchir A, Sangiuolo F, et al: Mandibuloacral dysplasia is caused by a mutation in LMNA-encoding lamin A/C. Am J Hum Genet 71:426–431, 2002 Ostlund C, Bonne G, Schwartz K, Worman HJ: Properties of lamin A mutants found in Emery–Dreifuss muscular dystrophy, cardiomyopathy and Dun- nigan-type partial lipodystrophy. J Cell Sci 114:4435–4445, 2001 Pendas AM, Zhou Z, Cadinanos J, et al: Defective prelamin A processing and muscular and adipocyte alterations in Zmpste24 metalloproteinase-de- ficient mice. Nat Genet 31:94–99, 2002 Sakaki M, Koike H, Takahashi N, Sasagawa N, Tomioka S, Arahata K, Ishiura S: Interaction between emerin and nuclear lamins. J Biochem (Tokyo) 129: 321–327, 2001 Schmidt WK, Tam A, Michaelis S: Reconstitution of the Ste24p-dependent N-terminal proteolytic step in yeast a-factor biogenesis. J Biol Chem 275: 6227–6233, 2000 Sermon K, Van Steirteghem A, Liebaers I: Preimplantation genetic diagnosis. Lancet 363:1633–1641, 2004 Sillevis Smitt JH, van Asperen CJ, Niessen CM, et al: Restrictive dermopathy. Report of 12 cases. Dutch Task Force on Genodermatology. Arch De- rmatol 134:577–579, 1998 Welsh KM, Smoller BR, Holbrook KA, Johnston K: Restrictive dermopathy. Re- port of two affected siblings and a review of the literature. Arch Dermatol 128:228–231, 1992 Witt DR, Hayden MR, Holbrook KA, Dale BA, Baldwin VJ, Taylor GP: Restrictive dermopathy: A newly recognized autosomal recessive skin dysplasia. Am J Med Genet 24:631–648, 1986 Worman HJ, Courvalin JC: How do mutations in lamins A and C cause disease? J Clin Invest 113:349–351, 2004 Zastrow MS, Vlcek S, Wilson KL: Proteins that bind A-type lamins: Integrating isolated clues. J Cell Sci 117:979–987, 2004 ZMPSTE24 MUTATIONS IN RESTRICTIVE DERMOPATHY 919125 : 5 NOVEMBER 2005