Entry - *107400 - SERPIN PEPTIDASE INHIBITOR, CLADE A, MEMBER 1; SERPINA1 - OMIM
 
* 107400

SERPIN PEPTIDASE INHIBITOR, CLADE A, MEMBER 1; SERPINA1


Alternative titles; symbols

ALPHA-1-ANTITRYPSIN; AAT
PROTEASE INHIBITOR 1; PI
PI1
ANTI-ELASTASE
ANTITRYPSIN


HGNC Approved Gene Symbol: SERPINA1

Cytogenetic location: 14q32.13     Genomic coordinates (GRCh38): 14:94,376,747-94,390,635 (from NCBI)


Gene-Phenotype Relationships
Location Phenotype Phenotype
MIM number
Inheritance Phenotype
mapping key
14q32.13 Emphysema due to AAT deficiency 613490 AR 3
Emphysema-cirrhosis, due to AAT deficiency 613490 AR 3
Hemorrhagic diathesis due to antithrombin Pittsburgh 613490 AR 3

TEXT

Description

The SERPINA1 gene encodes alpha-1-antitrypsin (AAT), also known as protease inhibitor (PI), a major plasma serine protease inhibitor. AAT complexes predominantly with elastase, but also with trypsin, chymotrypsin, thrombin, and bacterial proteases. The most important inhibitory action of AAT is that against neutrophil elastase (ELANE, or HLE; 130130), a protease that degrades elastin of the alveolar walls as well as other structural proteins of a variety of tissues (review by Cox, 2001).


Cloning and Expression

Kurachi et al. (1981) cloned a nearly full-length baboon AAT cDNA (approximately 1,352 bp) and a partial human AAT cDNA (approximately 306 bp). They found more than 96% homology between the cDNA and predicted amino acid sequences of AAT in the 2 species. Comparison of baboon AAT, human antithrombin III (107300), and chicken ovalbumin indicated about 30% homology of amino acid sequence.

Long et al. (1984) cloned a full-length human AAT cDNA from a liver cDNA library. Sequence analysis revealed a precursor molecule containing a 24-amino acid signal peptide and a mature protein of 394 amino acids. AAT is primarily synthesized in the liver.

Crystal (1990) noted that hepatocytes are the major source of AAT, but that the gene is also expressed in mononuclear phagocytes and neutrophils.


Gene Structure

Lai et al. (1983) showed that the AAT gene contains 3 introns in the peptide-coding region.

Long et al. (1984) found that the genomic length of the PI gene is 10.2 kb with a 1,434-bp coding region. The gene has 4 introns; exon 1, the 5-prime portion of exon 2, and the 3-prime portion of exon 5 are noncoding regions. The first intron, 5.3 kb long, contains a 143-amino acid open reading frame (which does not appear to be an actual protein coding region), an Alu family sequence, and a pseudotranscription initiation region.

Perlino et al. (1987) found that the AAT gene in macrophages is transcribed from a macrophage-specific promoter located about 2,000 bp upstream of the hepatocyte-specific promoter. Transcription from the 2 AAT promoters is mutually exclusive; the macrophage promoter is silent in hepatocytes and the hepatocyte promoter is silent in macrophages. In macrophages, 2 distinct mRNAs are generated by alternative splicing.

Hafeez et al. (1992) demonstrated that the AAT gene has 3 macrophage-specific transcriptional initiation sites upstream from a single hepatocyte-specific transcriptional initiation site. Macrophages use these sites during basal and modulated expression. Hepatoma cells use the hepatocyte-specific transcriptional initiation site during basal and modulated expression but also switch on transcription from the upstream macrophage transcriptional initiation sites during modulation by the acute phase mediator interleukin-6 (IL6; 147620).

Soutoglou and Talianidis (2002) analyzed the ordered recruitment of factors to the human alpha-1-antitrypsin promoter around the initial activation of the gene during enterocyte differentiation. They found that a complete preinitiation complex, including phosphorylated RNA Pol II (180660), was assembled at the promoter long before transcriptional activation. The histone acetyltransferases CBP (600140) and P/CAF (602303) were recruited subsequently, but local histone hyperacetylation was delayed. After transient recruitment of the human Brahma homolog BRM (600014), remodeling of the neighboring nucleosome coincided with transcription initiation. Soutoglou and Talianidis (2002) concluded that, at this promoter, chromatin reconfiguration is a defining step of the initiation process, acting after the assembly of the Pol II machinery.


Mapping

Lai et al. (1983) used a cloned AAT gene as a hybridization probe to analyze EcoRI-digested genomic DNA from different individuals and identified 2 distinct bands (9.6 kb and 8.5 kb long) in every case. Analysis using only intronic DNA as probe showed that the authentic gene resides in the 9.6-kb fragment. The 8.5-kb fragment was thought to contain a gene with close sequence homology to that of AAT.

By studying hybrids of mouse or rat hepatoma cells with human lymphocytes, Darlington et al. (1982) and Pearson et al. (1981) achieved direct assignment of the PI locus to chromosome 14. From study of 2 families with abnormalities of the long arm of chromosome 14, Cox et al. (1982) localized GM to 14q32.3 and PI to a more proximal position between 14q24.3 and 14q32.1. The immunoglobulin genes are in a chromosome region noted for its high frequency of breaks associated with chromosome rearrangement, occurring both spontaneously in cultured lymphocytes and in certain malignancies.

By in situ hybridization, Schroeder et al. (1985) narrowed the assignment of the PI locus to 14q31-q32. Turleau et al. (1984) studied a patient with an interstitial deletion of 14q and assigned the PI locus to 14q32.1 by exclusion mapping. In a similar patient with an interstitial deletion of 14q, Yamamoto et al. (1986) confirmed the assignment to 14q32.1. By the dosage principle, the level of alpha-1-antitrypsin in the patient was only about half of that in his parents and in controls.

Sefton et al. (1989) used pulsed field gel electrophoresis to demonstrate that the genes encoding alpha-1-antitrypsin and alpha-1-antichymotrypsin (AACT, SERPINA3; 107280) are approximately 220 kb apart and oriented in opposite directions.

Molecular studies of a ring chromosome 14 showed that the IGH and D14S1 loci were missing, whereas the PI locus was present (Keyeux et al., 1989). Thus, PI is proximal to the other 2 loci, a conclusion that was supported by much earlier data. A noncoding alpha-1-antitrypsin-like gene (PIL; 107410) is located 12 kb 3-prime of the AAT gene. Billingsley et al. (1989) found that this gene and the AAT and AACT genes are carried by a single 550-kb NarI fragment. Also see Billingsley et al. (1993).

By in situ hybridization, Ledbetter et al. (1987) localized the AAT locus to mouse chromosome 12.


Gene Function

Dycaico et al. (1988) established transgenic mouse lineages that carried the normal (M) (see 107400.0001) or mutant (Z) (107400.0011) alleles of the human AAT gene. All expressed the human protein in liver, cartilage, gut, kidneys, lymphoid macrophages, and thymus. The human M-allele protein was secreted normally into the serum. However, the human Z-allele protein accumulated in several cell types, particularly in hepatocytes, and was found in serum in concentrations 10 times lower than the M-allele protein. Mice in one lineage carrying the Z allele displayed significant runting in the neonatal period and had developed abnormalities in the liver with accumulation of human Z protein in diastase-resistant cytoplasmic globules that stained with periodic acid-Schiff reaction (PAS).

The major physiologic substrate of alpha-1-antitrypsin is elastase, particularly in the lower respiratory tract (Cox, 2001).

AAT is an acute-phase reactant in that serum levels are increased with inflammation, trauma, and pregnancy (Cox, 1989).

Alpha-1-Antitrypsin Deficiency

Deficiency of alpha-1-antitrypsin (613490) is primarily associated with the risk of emphysema and liver disease; see MOLECULAR GENETICS.

Role in Twinning

Lieberman et al. (1979) found an increased frequency of heterozygosity for antitrypsin deficiency in twins and parents of twins. They concluded that 'increased' fertility and twinning may be heterozygous advantages for antitrypsin deficiency. Clark and Martin (1982) found that the frequency of the S allele (107400.0013) in mothers of dizygotic twins (0.088) was double that in controls (0.044). The frequency of S in the parents of monozygotic twins and in fathers of DZ twins was no higher than in controls. Normal frequencies were observed for the Z allele (107400.0011). No fertility indices other than twinning itself were available. To study relationships between Pi types, fertility, and twinning, Boomsma et al. (1992) studied 90 DZ and 70 MZ Dutch twin pairs and their parents. They found that mothers of dizygotic twins had frequencies of the S and Z alleles that were 3 times higher than those in a control sample. Mothers of identical twins also had a higher frequency of S than controls. The S allele may thus both increase ovulation rate and enhance the success of multiple pregnancies.

Role in Human Immunodeficiency Virus-1 Infection

Bristow (2001) found that decreased human immunodeficiency virus (HIV) infectivity correlated significantly with decreased cell surface expression of leukocyte (neutrophil) elastase (HLE) on monocytes but not lymphocytes. Decreased levels of PI correlated with increased cell surface HLE expression and increased HIV infectivity.

Bristow et al. (2001) showed that decreased HIV viral load correlated with decreased circulating PI. Furthermore, asymptomatic patients manifested deficient levels of active PI. Bristow et al. (2001) noted that deficient levels of PI lead to degenerative lung diseases and suggested that preventing PI deficiency may prevent HIV-associated pathophysiology.

Using subclones of monocytic cell lines, Bristow et al. (2003) showed that HLE localized to the cell surface, but not granules, of HIV-1-permissive clones, and to the granules, but not the cell surface, of HIV-1-nonpermissive clones. Stimulation of nonpermissive clones with lipopolysaccharide and LBP (151990), followed by exogenous PI, induced cell surface HLE expression, resulting in susceptibility to HIV infection. PI appeared to promote HIV coreceptor colocalization with surface HLE, thus permitting HIV infectivity.

Shapiro et al. (2001) showed that, at physiologic concentrations, AAT and CE-2072, a synthetic inhibitor of neutrophil elastase and proteinase-3 (PRTN3; 177020), inhibited HIV-1 production in chronically infected monocytic cell lines, in fresh blood mononuclear cells infected after an activation step, and in permissive HeLa cells. EMSA analysis indicated that AAT suppressed activation of the HIV-1-inducing transcription factor NFKB (see 164011). In 5 individuals with the Z-type AAT mutation (glu342lys; 107400.0011), HIV-1 p24 antigen increased more than 6-fold in whole blood after infection with a monocyte-tropic HIV strain. In contrast, there was no significant increase in blood obtained from healthy volunteers.

By screening a peptide library generated from hemofiltrate, Munch et al. (2007) identified a 20-amino acid peptide from the C-proximal region of alpha-1-antitrypsin, designated virus-inhibitory peptide (VIRIP), as the most potent inhibitor of multiple HIV-1 strains, including those resistant to antiviral drugs. Changes in some VIRIP residues increased its antiviral potency 100-fold. VIRIP blocked HIV-1 entry by interacting with the virus gp41 fusion peptide. Munch et al. (2007) proposed that VIRIP may affect disease progression in HIV-1-infected individuals.

NET Inhibitory Peptides

Neonatal neutrophils fail to form neutrophil extracellular traps (NETs) due to circulating NET inhibitory peptides (NIPs), which are cleavage fragments of A1AT. Using immunofluorescence assays, Campbell et al. (2021) showed that human placenta from both term and preterm pregnancies secreted serine protease A1 (HTRA1; 602194) into fetal circulation. Plasma HTRA1 levels were reduced after delivery, and decreased HTRA1 plasma levels were associated with decreased levels of NIPs. Placental HTRA1 cleaved A1AT after amino acid 382 to generate a C-terminal cleavage fragment of A1AT, termed A1ATM383S-CF, that could inhibit NET formation in vitro. Through NET inhibition, A1ATM383S-CF decreased bacterial killing, but it maintained other key neutrophil activities in vitro. In vivo analysis with wildtype mice showed that mouse placenta also secreted Htra1, and placental Htra1 cleaved A1at to generate A1atM383S-CF and inhibit NET formation by neonatal neutrophils. Analysis with Htra1 -/- and wildtype mice revealed that inhibition of NET formation during experimental neonatal sepsis improved survival.


Nomenclature

Fagerhol (1968) suggested that the system of inherited AAT variants be called Pi for protease inhibitor. Cox (1978) reported the recommendations of a workshop on PI nomenclature.


Molecular Genetics

Many electrophoretic variants of serum alpha-1-antitrypsin have been described, beginning with those reported by Axelsson and Laurell (1965). Kueppers and Bearn (1967) studied an Italian family with multiple members heterozygous for an electrophoretic variant that could not be distinguished from that which Axelsson and Laurell (1965) found in a Swedish family.

About 30 variants of alpha-1-antitrypsin had been described by 1981 (Hug et al., 1981). The alleles were given symbols according to the relative electrophoretic mobility of the allele product.

Cox et al. (1987) studied RFLPs associated with the AAT gene. They gave information on extensive variability expressed by the polymorphic information content (PIC) as proposed by Botstein et al. (1980). PI types and M subtypes tended to be associated with specific RFLP haplotypes.

Nukiwa et al. (1988) indicated that approximately 75 AAT alleles had been identified at the protein and/or gene level.

Roychoudhury and Nei (1988) tabulated worldwide gene frequencies for allelic variants M (M1, M2, M3, M4), S, Z, F, I, and V. Cox (1989) and Crystal (1989) reviewed the variants, 'normal' and pathologic, of the PI gene.

'Normal' Alleles

Crystal (1989) listed 10 normal AAT alleles that had been sequenced (107400.0001-107400.0010).

Nukiwa et al. (1988) stated that the most common alleles are the 2 forms of M1, that with valine at position 213 (M1V; 107400.0002) and that with alanine at position 213 (M1A; 107400.0001).

'Risk' Alleles

Crystal (1989) divided AAT 'at risk' alleles into 'deficiency' alleles and 'null' alleles. He stated that except for the rare Pittsburgh allele (107400.0026), which is associated with a bleeding disorder, only those phenotypes comprising 2 'at risk' alleles place the individual at risk for development of disease. Alleles in the 'at risk' class are found almost exclusively among Caucasians of European descent, with the highest frequency in northern Europe. Blacks and Asians are rarely affected.

The most common AAT deficiency allele is the Z allele (glu342-to lys; 107400.0011), which occurs on an M1A (ala213; 107400.0001) haplotype background (Nukiwa et al., 1986). The homozygous ZZ phenotype is associated with a high risk of both emphysema and liver disease. The Z allele reaches polymorphic frequencies in Caucasians and is rare or absent in Asians and blacks (DeCroo et al., 1991; Hutchison, 1998).

Clark et al. (1982) reported the cases of 2 brothers with Weber-Christian panniculitis and the AAT Z phenotype. A younger brother had the Z phenotype without Weber-Christian disease. Along with several earlier reported cases, these observations establish a relationship.

Another common AAT deficiency allele is the S allele (glu264-to-val; 107400.0013), which occurs on an M1V (val213; 107400.0002) haplotype background. Pi*S homozygotes are at no risk of emphysema, but compound heterozygotes with a Z or a null allele have a mildly increased risk (Curiel et al., 1989). The S allele reaches polymorphic frequencies in Caucasians and is rare or absent in Asians and blacks. It is not associated with liver disease.

Other rare deficiency AAT alleles may result in increased risk for both liver and lung disease (e.g., Pi M(Malton); 107400.0012) or only emphysema (e.g., Pi M(Procida); 107400.0016). Some of the rare deficiency alleles have been found in Japanese (e.g., Pi S(Iiyama); 107400.0039).

Null AAT alleles are rare but have been found in all populations. Garver et al. (1986) investigated the molecular basis of the Pi null-null AAT phenotype. The gene appeared to be intact without discernible deletion or other structural abnormality, yet there was no detectable mRNA produced. The 5-prime promoter region also appeared to be normal. No evidence of hypermethylation of cytosine nucleotides in the promoter region was detected. The defect may be comparable to that in some forms of thalassemia in which a change, at a splicing site, for example, may lead to greatly reduced mRNA production. The null-null phenotype is accompanied by emphysema as is the ZZ and SZ phenotypes but an important difference is that cirrhosis and liver disease do not occur with the null-null phenotype; there is no abnormal antitrypsin produced that is excreted with difficulty from the cells of synthesis.

Nukiwa et al. (1987) identified a null form of alpha-1-antitrypsin resulting from a frameshift causing a stop codon to be formed approximately 44% from the N terminus of the precursor protein (Null(Granite Falls); 107400.0020). Although the molecular basis of antitrypsin deficiency was quite different from that in the Z haplotype, the phenotypic consequences were similar: severe deficiency associated with high risk of emphysema.

Seixas et al. (2002) reported 2 null alleles of the PI gene in Portuguese patients with emphysema. These alleles were associated with total lack of circulating protein as indicated by the absence of isoelectric focusing banding patterns. One of the alleles, designated Q0(Ourem), was identical to Q0(Mattawa) on an M3 normal background (107400.0022). The second allele, Q0(Porto), had a novel mutation which restricted mononuclear phagocyte transcripts to mRNA species resulting from the direct splice of exon IA to exon II. The absence of this normal splice alternative in the liver, where PI is primarily synthesized, provided a basis for the pathogenic effects of this mutation.

PI Pittsburgh

The PI Pittsburgh allele (met358-to-arg; 107400.0026), which occurs at the AAT active site, is an example of a mutation leading to altered function of the gene product. AAT becomes a potent inhibitor of thrombin and factor XI rather than of elastase. The mutation results in a bleeding disorder (Lewis et al., 1978; Owen et al., 1983).

SERPINA1 Haplotypes Associated with Chronic Obstructive Pulmonary Disease

While cigarette smoking is a major cause of COPD (see 606963), only 15% of smokers develop the disease, indicating major genetic influences. The most widely recognized candidate gene in COPD is SERPINA1, although it has been suggested that SERPINA3 (107280) may also play a role. Chappell et al. (2006) identified 15 single-nucleotide polymorphism (SNP) haplotype tags from high-density SNP maps of the 2 genes and evaluated these SNPs in the largest case-control genetic study of COPD conducted to that time. For SERPINA1, 6 newly identified haplotypes with a common backbone of 5 SNPs were found to increase the risk of disease by 6- to 50-fold, the highest risk of COPD that had been reported. In contrast, no haplotype associations for SERPINA3 were identified.


Population Genetics

DeCroo et al. (1991) studied the frequency of alpha-1-antitrypsin alleles in US whites, US blacks, and African blacks (living in Nigeria). While the PI*S allele was present at a polymorphic level in US whites, it was present only sporadically in US blacks and completely absent in African blacks. The PI*Z allele was not detected in the black populations tested. DeCroo et al. (1991) used the PI allele frequency data to calculate white admixture in US blacks. The average white admixture estimate in US blacks, based on all PI alleles, was about 13%. This value was about 24% when only the S and Z alleles were used.

Studies of the distribution of the S and Z alleles in Europe demonstrated that they occur mainly among those of European stock. Hutchison (1998) found that the frequency of the gene for PiZ is highest on the northwestern seaboard of the continent and that the mutation seems to have arisen in southern Scandinavia. The distribution of PiS is quite different: the gene frequency is highest in the Iberian peninsula and the mutation is likely to have arisen in that region.

By means of a metaanalysis of 43 studies, Blanco et al. (2001) analyzed the distribution of the PI*S and PI*Z alleles in countries outside Europe and compared them with data from Europe.


Animal Model

The pallid (pa) (604310) mouse develops emphysema late in life. Martorana et al. (1993) demonstrated that pallid mice have markedly reduced levels of serum alpha-1-antitrypsin associated with severe deficiency in serum elastase inhibitory capacity. However, they have normal alpha-1-antitrypsin mRNA levels in the liver.

Green et al. (2003) showed that Drosophila 'necrotic' (nec) mutations can mimic alpha-1-antitrypsin deficiency. They identified 2 nec mutations homologous to an antithrombin point mutation that is responsible for neonatal thrombosis. Transgenic flies carrying an amino acid substitution equivalent to that found in S(Iiyama) variant antitrypsin (107400.0039) failed to complement nec-null mutations and demonstrated a dominant temperature-dependent inactivation of the wildtype nec allele. Green et al. (2003) concluded that the Drosophila nec system can be used as a powerful system to study serpin polymerization in vivo.

Van Pel et al. (2006) reported that IL8 (146930)- and GCSF (CSF3; 138970)-induced mobilization of hematopoietic stem cells (HSCs) and hematopoietic progenitor cells (HPCs) in mice was completely inhibited by total body irradiation (TBI). They found that TBI increased expression of Serpina1 mRNA and protein, which inhibited elastase activity. Inhibition of HSC/HPC mobilization in irradiated mice could be reversed by anti-Serpina1. Furthermore, injection of Serpina1, but not heat-inactivated Serpina1, prior to administration of IL8 inhibited HSC/HPC mobilization. Van Pel et al. (2006) concluded that low-dose TBI induces Serpina1 in bone marrow and inhibits HSC/HPC mobilization, and they hypothesized that cytokine-induced HSC/HPC mobilization is determined by a critical balance between serine proteases and their inhibitors.

Kamimoto et al. (2006) bred the PiZ mouse onto the GFP-LC3 background to monitor autophagy.

Hidvegi et al. (2010) demonstrated that the autophagy-enhancing drug carbamazepine decreased the hepatic load of mutant alpha-1-antitrypsin Z (ATZ) protein and hepatic fibrosis in a mouse model of AAT deficiency-associated liver disease. The mouse used was the PiZ mouse, developed by Dycaico et al. (1988), in which the human ATZ gene is a transgene. Although the PiZ mouse has normal circulating levels of endogenous murine alpha-1-antitrypsin, it is a robust model of liver disease associated with AAT deficiency, as characterized by intrahepatocytic ATZ-containing globules, inflammation, and increased regenerative activity, dysplasia, and fibrosis. Hidvegi et al. (2010) concluded that their results in this animal model provided a basis for testing carbamazepine, which has an extensive clinical safety profile in patients with AAT deficiency (613490) and also provided proof of principle for therapeutic use of autophagy enhancers.


History

The polymorphism of prealbumin described by Fagerhol and Braend (1965) was shown by Fagerhol and Laurell (1967) to be the same as the alpha-1-antitrypsin polymorphism.

A possible heterogeneity in recombination frequency between Pi variants believed to be allelic was reported by Gedde-Dahl et al. (1972): Pi(Z) had less recombination with Gm than Pi(non-Z). Gedde-Dahl et al. (1975) gave further data on the Gm-Pi linkage. They considered heterogeneity of recombination fraction among males of different Pi types to be likely. The major difference seemed to be between the Pi(Z) and other alleles. Possible explanations included a chromosomal deletion, inversion or locus regulating recombination in linkage disequilibrium with the Pi locus. Gedde-Dahl et al. (1981) showed that the allele-specific heterogeneity of Gm-Pi linkage is attributable to 'reduced' recombination in Z-allele heterozygotes. They found an equal sex ratio for Pi 'non-Z' variants, as opposed to a 1:2 male-female ratio for 'Z' families. The location of Gm and Pi on 6p was excluded by Bender et al. (1979).

Babron et al. (1990) confirmed a previous finding that the presence of the Pi Z allele tends to decrease the recombination rate between the GM (147100) and PI loci. This decrease appeared to be similar in both sexes and not unique in males as previously noted. The results suggested a possible linkage disequilibrium between the Pi Z allele and a large inversion between the GM and PI loci.

Carrell (1986) cited evidence for the existence of 2 genes coding for alpha-1-antitrypsin, although the plasma findings were compatible with expression of the alleles at a single locus.


ALLELIC VARIANTS ( 40 Selected Examples):

.0001 PI M1-ALA213

PI, M1A
SERPINA1, ALA213
   RCV000019553...

M1A, a normal variant, is believed to be the 'oldest' human PI allele, with the other common normal alleles M1V (107400.0002), M2 (107400.0003), and M3 (107400.0004) derived from M1A by single base substitutions. M2 is derived from M3; it has the same amino acid difference that distinguishes M3 from M1V but a second substitution in addition. The 4 common normal alleles are considered the 'base' from which all the other alleles are derived (see Fig. 4 in Crystal, 1989). The M1A allele has a frequency of 0.20-0.23 in US Caucasians.


.0002 PI M1-VAL213

PI, M1V
SERPINA1, ALA213
   RCV000019555...

This normal allele has a frequency of 0.44-0.49 in US Caucasians.


.0003 PI M2

SERPINA1, ARG101HIS ON M3
  
RCV000019557...

M2, which has a frequency of 0.10-0.11 in US Caucasians (Cox, 1989), was studied by Nukiwa et al. (1988), who found that its coding exons are identical to those of the more frequent form of M1 (val213) except for 2 bases: a change in codon 101 from CGT to CAT, leading to an amino acid change of arginine to histidine; and a change in codon 376 from GAA to GAC, resulting in an amino acid change from glutamic acid to aspartic acid. Since 2 mutations separate these 2 common alleles, Nukiwa et al. (1988) suggested that another AAT variant (presumably M3) was an intermediate in their evolution.


.0004 PI M3

SERPINA1, GLU376ASP ON M1V
   RCV000019558...

This normal variant allele has a frequency of 0.14-0.19 among US Caucasians. Graham et al. (1990) identified a single nucleotide difference between M1 (val213) and M3: a transversion in codon 376 from GAA(glu) to GAC(asp).


.0005 PI M4

SERPINA1, ARG101HIS ON M1V
   RCV000019557...

M4, an uncommon normal allele, is likely derived by single substitution from M1V; however, it has the same mutation that changed M2 to M3, and thus it is possible that M4 derived from M3 (or vice versa).


.0006 PI B(ALHAMBRA)

SERPINA1, ASP-LYS
   RCV000019560

Yoshida et al. (1979) found 2 amino acid substitutions in the rare antitrypsin variant PiB Alhambra. One substitution was asp for lys at an unknown location (Crystal, 1989).


.0007 PI F

SERPINA1, ARG223CYS ON M1V
  
RCV000019561...

This rare 'normal' allele has a CGT-to-TGT change in codon 223 (Crystal, 1989; Okayama et al., 1991).


.0008 PI P(ST. ALBANS)

SERPINA1, ASP341ASN ON M1V
  
RCV000019562...

In addition to a GAC-to-AAC change in codon 341, this rare 'normal' allele has a 'silent' asp256 (GAT)-to-asp256 (GAC) change. The rare P-family of AAT variants is defined by the position of migration of the protein on isoelectric focusing (IEF) of serum between the common M and S variants. The P(St. Albans) allele is associated with normal serum levels of AAT, whereas the P(Lowell) allele (107400.0019) is associated with reduced levels. Holmes et al. (1990) described the DNA change underlying both of these variants.


.0009 PI X

SERPINA1, GLU204LYS ON M1V
  
RCV000019563...

This rare 'normal' allele has been sequenced only at the level of the protein (Crystal, 1989).


.0010 PI CHRISTCHURCH

SERPINA1, GLU363LYS
  
RCV000019564...

Brennan and Carrell (1986) characterized antitrypsin Christchurch, which shows a substitution of lysine for glutamic acid at position 363. Although electrophoretic mobility of the mutant protein was abnormal, no functional abnormality of the protein was detected. The base PI allele, M1A or M1V, is unknown (Crystal, 1989).


.0011 PI Z

SERPINA1, GLU342LYS ON M1A
  
RCV000019567...

This is the most frequent allele leading to a high risk of emphysema (and liver disease) in the homozygote; the allele frequency is 0.01-0.02 in US Caucasians (Crystal, 1989). Nukiwa et al. (1986) demonstrated the val213-to-ala substitution (here symbolized M1A) in PI*Z in addition to the disease-producing glu342-to-lys mutation. Ala213 was found in all of 40 Z haplotypes, using synthetic oligonucleotide gene probes directed toward the mutated exon 3 sequences in the Z gene. Furthermore, the exon 3 mutation eliminated a BstEII restriction endonuclease site, allowing rapid identification of the change in genomic DNA. Surprisingly, only 23% of the M1 haplotypes were found to be BstEII site negative. The new form of M1, i.e., M1(ala213), is identical to M1 but has an isoelectric focusing 'silent' amino acid substitution. M1 has a frequency of 68 to 76%; M2, 14 to 20%; and M3, 10 to 12%. The Z gene represents 1 to 2% of all alpha-1-antitrypsin haplotypes.

Using 2 genomic probes extending into the 5-prime and 3-prime flanking regions, respectively, Cox et al. (1985) identified 8 polymorphic restriction sites for the PI gene. Extensive linkage disequilibrium was found with the PI Z allele throughout the probe region, but not with the normal PI M allele. The Z allele occurred mainly with one haplotype, indicating a single, relatively recent origin in Caucasians. This was an individual who lived in northern Europe some 6,000 years ago. Since then, the variant has spread through Europe with a frequency gradient extending from north to south: 5% of Scandinavians, 4% of Britons, 1 to 2% of southern Europeans, and 3% of the heterogeneous white population in the United States are MZ heterozygotes. Curiously, there is a reciprocal distribution of the S variant form: 10% in southern Europe to 5% in the north. As a general rule then, 1 in 10 persons of European origin will be heterozygous for either the S or Z variant, i.e., MZ or MS (Carrell, 1986). Kawakami et al. (1981) cited 2 studies in which no Pi Z was found among 965 healthy Japanese and 183 Japanese with pulmonary diseases. This is to be compared with a frequency of 1.6% for Pi Z among Norwegians.

Crystallographic analysis of alpha-1-antitrypsin predicts that in the normal protein a negatively charged glu342 is adjacent to a positively charged lys290. Thus, the glu342-to-lys Z mutation causes the loss of a normal salt bridge, resulting in intracellular aggregation of the Z molecule. Brantly et al. (1988) predicted that a second mutation that changed the positively charged lys290 to a negatively charged glu290 would correct the secretion defect. They demonstrated that such was the case: when the second mutation was added to the Z-type cDNA, the resulting gene directed the synthesis and secretion of AAT similar to that directed by the normal AAT cDNA in an in vitro eukaryotic expression system. In general it may be possible to correct human hereditary disease by inserting an additional mutation in the gene.

By analyzing nonrecombinant SNPs of 21 Latvian and 65 Swedish heterozygous and homozygous PI Z allele carriers and 113 healthy Latvian controls, Lace et al. (2008) estimated the age of the PI Z mutation to be 2,902 years in Latvia and 2,362 years in Sweden. The SNPs showed a high degree of similarity between the 2 populations, indicating a common ancestor.

Approximately 3 to 5% of patients with cystic fibrosis (CF; 219700) develop severe liver disease defined as cirrhosis with portal hypertension. Bartlett et al. (2009) performed a 2-stage case control study enrolling patients with CF and severe liver disease with portal hypertension from 63 CF centers in the United States as well as 32 in Canada and 18 outside of North America. In the first stage, 124 patients with CF and severe liver disease, enrolled between January 1999 and December 2004, and 843 control patients without CF-related liver disease (all assessed at greater than 15 years of age) were studied by genotyping 9 polymorphisms in 5 genes previously studied as modifiers of liver disease in CF. In the second stage, the 2 genes that were positive from the first stage were tested in an additional 136 patients with CF-related liver disease enrolled between January 2005 and February 2007, and in 1,088 with no CF-related liver disease. The combined analysis of the initial and replication studies by logistic regression showed CF-related liver disease to be associated with the SERPINA1 Z allele (odds ratio = 5.04; 95% confidence interval 2.88-8.83; p = 1.5 x 10(-8)). Bartlett et al. (2009) concluded that the SERPINA1 Z allele is a risk factor for liver disease in CF. Patients carrying the Z allele are at greater risk (odds ratio = approximately 5) of developing severe liver disease with portal hypertension.

Using several markers of ER stress response, Kelly et al. (2009) found that expression of AAT with the Z mutation, which they called ZAAT, resulted in ER stress in transfected HepG2 cells. Coexpression of the ER stress response selenoprotein SEPS1 (607918) relieved ER stress caused by ZAAT expression or by exposure to tunicamycin, a known ER stressor. Supplementation of cells with selenium augmented the activity of SEPS1. Selenium supplementation also increased endogenous SEPS1 expression and reduced ER stress.


.0012 PI M(MALTON)

SERPINA1, PHE52DEL ON M2
  
RCV000019568...

Liver disease, as well as emphysema, has been described in patients with the rare PI*M(Malton) allele. Fraizer et al. (1989) studied the molecular defect in M(Malton), a deficiency allele which, like the Z allele, is associated with hepatocyte inclusions and impairs secretion. They found that the M(Malton) allele contains a deletion of the codon for 1 of the 2 adjacent phenylalanine residues (amino acid 51 or 52 of the mature protein). Judging from the haplotype data, the M(Malton) mutation must have derived from the normal M2 allele. Deletion of the 1 amino acid would be expected to shorten 1 strand of the beta-sheet, B6, apparently preventing normal processing and secretion. Curiel et al. (1989) also showed that the M(Malton) allele differs from the normal M2 allele by deletion of the entire codon (TTC) for residue phe52. They demonstrated abnormal intracellular accumulation of newly synthesized AAT protein in a homozygote who also showed, on liver biopsy, inflammation, mild fibrosis, and intrahepatocyte accumulation of the protein. Furthermore, Curiel et al. (1989) showed by retroviral gene transfer of AAT cDNA with the M(Malton) phe52 deletion into murine cells that abnormal accumulation of the newly synthesized protein occurred. This provides further evidence that abnormal intrahepatocyte AAT accumulation is responsible for the liver injury. By means of gene amplification and direct DNA sequencing, Graham et al. (1989) identified the same mutation, pointing out that it could be either phenylalanine-51 or phenylalanine-52 that is deleted.


.0013 PI S

SERPINA1, GLU264VAL ON M1V
  
RCV000019569...

Owen and Carrell (1976) and Yoshida et al. (1977) found substitution of valine for glutamic acid at position 264 in the S variant of alpha-1-antitrypsin. See Long et al. (1984).

Curiel et al. (1989) concluded that the S-type AAT protein is degraded intracellularly before secretion. PI*S homozygotes are at no risk of emphysema, but compound heterozygotes with Z or a null allele have a mildly increased risk. Because of the high frequency of the PI*S allele (0.02-0.04 in US Caucasians), such compound heterozygotes are relatively frequent.


.0014 PI M(HEERLEN)

SERPINA1, PRO369LEU ON M1A
  
RCV000019565...

Hofker et al. (1989) demonstrated the molecular defect in the PI gene of a patient with a serum level of only 5 mg/100 ml and a PI M-like phenotype, designated PI M(Heerlen). They demonstrated a substitution of leucine for proline at codon 369, which resulted from a C-to-T mutation in exon 5. Otherwise the nucleotide sequence of the exons, intron/exon junctions, and a part of the promoter region was similar to that of a PI M1(ala213) gene. Kalsheker et al. (1992) described a family with Pi M(Heerlen) and commented on the difficulties of diagnosis of rare PI (null) or Q0 variants.


.0015 PI M(MINERAL SPRINGS)

SERPINA1, GLY67GLU ON M1A
  
RCV000019566...

This mutation, which causes AAT deficiency and emphysema, is unique among antitrypsin mutations in that it was observed in a black family, whereas most mutations causing AAT deficiency are confined to Caucasian populations of European descent. The index case was homozygous. A GGG-to-GAG change in codon 67 led to substitution of glutamic acid for glycine (Curiel et al., 1990). Curiel et al. (1990) showed that this mutation caused reduced AAT secretion on the basis of aberrant posttranslational biosynthesis by a mechanism distinct from that associated with the Z allele, whereby intracellular aggregation of the mutant protein is responsible for the secretory defect. Furthermore, the M(Mineral Springs) mutation markedly affected the ability of the protein that did reach the circulation to inhibit neutrophil elastase. Homozygotes have a high risk of emphysema (Crystal, 1989).


.0016 PI M(PROCIDA)

SERPINA1, LEU41PRO ON M1V
  
RCV000019571...

Takahashi et al. (1988) showed that M(Procida) has a substitution of proline for leucine at position 41, resulting from a change of codon CTG to CCG. The rare mutant protein shows somewhat reduced catalytic activity; its concentration is low in plasma, apparently because of instability and resulting intracellular degradation before secretion. Homozygotes have a high risk of emphysema (Crystal, 1989).


.0017 PI M(NICHINAN)

SERPINA1, PHE52DEL AND GLY148ARG
  
RCV000019568...

Nakamura et al. (1980) found this variant in a 42-year-old Japanese woman with neither pulmonary emphysema nor liver dysfunction. She was the product of a consanguineous marriage. Radial immunodiffusion assay showed a low level of AAT in serum (17.9 mg/dl as compared to the normal range of 190-280 mg/dl). Aggregation of AAT molecules was demonstrated histologically in hepatocytes, indicating profound reduction in the secretion of the protein. Serum AAT levels in the members of the family demonstrated that the proband was homozygous for the M(Nichinan) allele. Matsunaga et al. (1990) demonstrated that the M(Nichinan) gene is identical with the M1(val213) gene except for 2 changes: a TTC trinucleotide deletion in the codon for phenylalanine-52 and a G-A substitution by which the normal gly148(GGG) became arg148(AGG). Matsunaga et al. (1990) suggested that the gly148-to-arg change is unlikely to be the cause of the AAT deficiency because arg (not gly) is located at the corresponding position of the protein C inhibitor which belongs to the same family of serine protease. On the other hand, Matsunaga et al. (1990) suggested that deletion of phenylalanine-52 may cause the newly synthesized AAT protein to aggregate, resulting in serum AAT deficiency. They suggested that the gly148-to-arg substitution reflects the vulnerability of a CpG dinucleotide to mutation. They pointed to a number of other variant forms of AAT that were probably generated through a C-T transition. Indeed, the Z and M1(val213) genes were generated from the M1(ala213) gene by the C-T transition at the CpG dinucleotide on the antisense and the sense strands, respectively. The M2 gene was generated from the M3 gene by the same mechanism.


.0018 PI I

SERPINA1, ARG39CYS ON M1V
  
RCV000019575...

By gene amplification and direct DNA sequencing, Graham et al. (1989) identified this mutation, CGC to TGC, in a compound heterozygote. Homozygotes are at no risk of emphysema, but compound heterozygotes with Z or a null allele have a mildly increased risk (Crystal, 1989). In 1 individual and 3 independent families, Seri et al. (1992) confirmed that the I variant resulted from a CGC (arg)-to-TGC (cys) transition at codon 39 within exon 2.


.0019 PI P(LOWELL)

PI NULL(CARDIFF)
PI Q0(CARDIFF)
SERPINA1, ASP256VAL ON M1V
  
RCV000019576...

Faber et al. (1989) demonstrated that this rare allele, a cause of deficiency of alpha-1-antitrypsin, results from an A-to-T transversion in exon 3 of the gene. As a result, GAT (aspartic acid at residue 256) is converted to GTT (valine at that position). The same change was found in a total of 4 families.

By gene amplification and direct DNA sequencing, Graham et al. (1989) identified the same mutation in a variant they called Null(Cardiff). According to the tabulation by Crystal (1989), homozygotes have no risk for emphysema, but compound heterozygotes with a Z or null allele have a mildly increased risk.

By retroviral insertion of the P(Lowell) cDNA into the genome of NIH-3T3 fibroblasts, Holmes et al. (1990) demonstrated a pattern of biosynthesis of AAT consistent with the intracellular degradation of newly synthesized protein. Because serum AAT deficiency associated with other mutations resulting from intracellular degradation of the protein can be overcome by administration of estrogenlike drugs, Holmes et al. (1990) administered tamoxifen to a subject with the P(Lowell)/Z phenotype and demonstrated a 48% rise in AAT serum levels over a 5-month period, from below the threshold for protection from emphysema to a value above that threshold. Seri et al. (1992) confirmed the nature of the mutation in P(Lowell).

Hildesheim et al. (1993) demonstrated that P(Duarte) (107400.0037) has the same mutation as that in P(Lowell) but that it is on a background of the normal M4 allele (R101H; 107400.0005). Hildesheim et al. (1993) pointed out that this is an example of genetic diversity resulting from a limited repertoire of mutations on different common allelic backgrounds--a combinatorial basis for genetic diversity. A similar example is the occurrence of Creutzfeldt-Jakob disease and fatal familial insomnia as a result of the same mutation, depending on the nature of a nucleotide polymorphism at another site in the prion protein gene (PRNP; 176640.0010).


.0020 PI NULL(GRANITE FALLS)

PI Q0(GRANITE FALLS)
SERPINA1, TYR160TER ON M1A
  
RCV000019579...

The gene shows deletion of the third nucleotide in the tyr160 codon TAC, causing a frameshift with new stop codon TAG at position 160 (Nukiwa et al., 1987). Emphysema is associated with homozygosity.


.0021 PI NULL(BELLINGHAM)

PI Q0(BELLINGHAM)
SERPINA1, LYS217TER ON M1V
  
RCV000019581...

By cloning and sequencing the Null(Bellingham) gene (which in homozygous state is associated with early-onset emphysema), Satoh et al. (1988) demonstrated that the promoter region, coding exons, and all exon-intron junctions are normal except for a single base substitution in exon 3, which causes the normal lys217 (AAG) to become a stop codon (TAG).


.0022 PI NULL(MATTAWA)

SERPINA1, LEU353PHE ON M1V
  
RCV000019583

Cox and Levison (1988) reported a family in which several members manifested no detectable plasma alpha-1-antitrypsin (613490), indicating a 'null' AAT allele, which the authors designated Null Mattawa (QO-Mattawa). Curiel et al. (1989) studied 2 affected sisters in this family and found that they were compound heterozygous for 2 mutations in the AAT gene: Null(Bellingham) (107400.0021) and Null(Mattawa). Sequencing of exons 1c-5 and all exon-intron junctions of the Null(Mattawa) gene demonstrated that it was identical to the common normal M1(val213) gene except for the insertion of a single nucleotide within the coding region of exon 5, causing a 3-prime frameshift with generation of a premature stop signal at position 376. Monocytes were shown to have an mRNA transcript of normal size, and in vitro translation showed that the mRNA was translated at a normal rate but produced a truncated antitrypsin protein. Additionally, retroviral transfer of the cDNA to murine fibroblasts demonstrated no detectable intracellular or secreted protein despite the presence of Null(Mattawa) mRNA. Thus, the molecular pathophysiology of Null(Mattawa) is probably manifested at a posttranslational level. This allele is associated with high risk of emphysema.


.0023 PI NULL(PROCIDA)

PI NULL(ISOLA DI PROCIDA)
PI Q0(PROCIDA)
SERPINA1, 17-KB DEL
   RCV000019584...

Of the 5 previously known representatives of the 'null' group of AAT-deficient alleles (i.e., genes incapable of producing AAT protein detectable in serum) evaluated at the gene level, all had stop codons in coding exons. Cloning and mapping of the Null(Isola di Procida) gene demonstrated deletion of a 17-kb fragment that included exons 2-5 of the AAT structural gene (Takahashi and Crystal, 1990). Sequence analysis showed a 7-bp repeat sequence both 5-prime to the deletion and at the 3-prime end of the deletion, suggesting that the mechanism of the deletion may have been a slipped mispairing. This mutation, which at first was called Null(Procida), was found in heterozygous state with the M(Procida) allele (107400.0016) reported by Takahashi et al. (1988). To avoid confusion with M(Procida), Null(Procida) was renamed Null(Isola di Procida). This mutation is associated with high risk of emphysema.


.0024 PI NULL(HONG KONG 1)

PI Q0(HONG KONG 1)
SERPINA1, 2-BP DEL, FS334TER
  
RCV000019587...

Deletion of TC from CTC codon 318 for leucine causes frameshift with stop codon TAA at position 334. Homozygosity for this allele, like other null alleles, predisposes to early-onset emphysema. See Sifers et al. (1988). This variant was initially called Null(Hong Kong) but later Null(Hong Kong-1) because a second null allele called Null(Hong Kong-2) (107400.0034) was identified in the same individual by haplotype analysis (Fraizer et al., 1990).


.0025 PI NULL(BOLTON)

PI Q0(BOLTON)
SERPINA1, 1-BP DEL
  
RCV000019589...

Fraizer et al. (1989) observed a unique PI null allele. By cloning and sequencing the allele, they demonstrated deletion of a single cytosine residue (the third C in the CCC codon 362 for proline) near the active site of alpha-1-antitrypsin in exon 5 resulting in a frameshift which caused an in-frame stop codon downstream of the deletion. The stop codon led to premature termination of protein translation at amino acid 373, resulting in a truncated protein. PI Q0(Bolton) was observed in combination with PI*M(Malton) in 2 compound heterozygotes. The allele carries a high risk of emphysema.


.0026 PI PITTSBURGH

'ANTITHROMBIN' PITTSBURGH
SERPINA1, MET358ARG
  
RCV000019591...

This structure mutation in the PI gene alters its function such that it becomes an antithrombin and leads to a bleeding disorder. Alpha-1-antitrypsin and antithrombin III (107300) have a similar structure reflecting origin from a common ancestral protein some 500 million years ago. Both are inhibitors of proteolytic enzymes but have different specificities. Alpha-1-antitrypsin protects the body against released elastase, whereas AT III controls coagulation by inhibiting thrombin and other activated coagulation factors. Owen et al. (1983) described a mutation of alpha-1-antitrypsin that converts it to an antithrombin. Whereas synthesis of alpha-1-antitrypsin increases in response to trauma, AT III remains at a constant plasma concentration and requires activation by heparin. The antithrombin activity of the mutant alpha-1-antitrypsin was independent of heparin but its synthesis was stimulated by trauma. The patient was a 14-year-old boy who died in 1981 with a huge hematoma of his leg and abdomen. This was the last of a lifelong series of bleeding episodes occurring after trauma and requiring hospitalization on more than 50 occasions. Lewis et al. (1978) described the clinical picture and identified a variant 'antithrombin' which they called antithrombin Pittsburgh. It had, however, the electrophoretic and antigenic characteristics of a variant alpha-1-antitrypsin. Owen et al. (1983) showed that the variant protein has arginine at position 358, replacing the normal methionine. This finding indicated that the reactive center of alpha-1-antitrypsin is methionine 358, which acts as a 'bait' for elastase, just as the normal reactive center of AT III is arginine-393, which acts as a bait for thrombin. Neutrophils augment tissue proteolysis by the oxidative inactivation of the methionine at the reactive center of alpha-1-antitrypsin. Scott et al. (1986) and Schapira et al. (1986) found that recombinant AAT-Pittsburgh (met358-to-arg) is a potent inhibitor of plasma kallikrein and activated factor XII fragment, although it has lost its anti-elastase activity. They suggested it might have therapeutic potential in hereditary angioedema or septic shock. Vidaud et al. (1992) demonstrated that a G-to-T transversion at nucleotide 10038 is responsible for the substitution of arg for met, which converts alpha-1-antitrypsin into an arg-ser protease inhibitor (serpin) that inhibits thrombin and factor Xa more effectively than antithrombin III. They observed a 15-year-old boy who surprisingly had no bleeding history. They suggested that a large decrease in protein C concentration may account for the mild or absent bleeding tendency. The deficiency of protein C in turn was attributed to deleterious effect of the abnormal inhibitor on both intracellular processing and catabolism of protein C. In later studies, Emmerich et al. (1995) suggested that strong affinity of the mutant AAT for protein C leads in the patient of Vidaud et al. (1992) to an increased turnover and thus to a low circulating level of protein C. They proposed that in the presence of the Pittsburgh mutant protein C can be activated and is abnormally rapidly cleared. The resultant relative lack of protein C anticoagulant function may ameliorate the bleeding diathesis expected to be associated with the Pittsburgh mutation.

Wilkie (1994) discussed the molecular basis of genetic dominance and provided a useful table. He indicated altered substrate specificity as one mechanism and antithrombin Pittsburgh as a specific example.


.0027 PI V(MUNICH)

SERPINA1, ASP2ALA ON M1V
  
RCV000019593...

In an alpha-1-antitrypsin variant called V(Munich) because the major fraction focused in the 'V' region of the isoelectric focusing gel, Holmes et al. (1990) found that the molecule differs from that of the common M1V allele by a single nucleotide substitution of cytosine for adenosine, with the resultant amino acid change asp2 to ala; the codon change is GAT to GCT.


.0028 PI Z(AUGSBURG)

PI Z(TUN)
SERPINA1, GLU342LYS ON M2
   RCV000019567...

Using isoelectric focusing with a narrow pH gradient, Weidinger et al. (1985) recognized a rare deficient PI variant, which they called PI Z(Augsburg). To their surprise, Faber et al. (1990) found that the sequence of the Z(Augsburg) gene showed the common PI*Z mutation (M1 glu342 GAG to Z lys342 AAG) which occurred, however, in an M2 ancestral gene. Previous findings indicated that the Z mutation had always been derived from an M1 ala213 background gene. Whitehouse et al. (1989) studied 2 sibs with mild liver abnormality who were found to be compound heterozygotes for the classical PI*Z allele and an allele that they called PI*Z(Tun). The Z(Tun) protein appeared to be deficient in the plasma to about the same degree as the Z protein. They found that the mutation was precisely the same as that in the Z allele, namely, a G-to-A transition at codon 342 resulting in the substitution of lysine for glutamic acid; however, the Z(Tun) mutation had occurred on an M2-like haplotype background rather than the M1A background. Because of its association with a unique DNA haplotype and the gene frequency estimates in populations of European origin, the Z mutation is thought to have occurred only once, about 6,000 years ago, in a North European person. The Z gene is very rare among other ethnic groups.


.0029 PI W(BETHESDA)

SERPINA1, ALA336THR ON M1A
  
RCV000019596...

This variant allele, which is associated with increased risk of emphysema and liver disease, has a mutation in exon 5 where codon 336 is changed from GCT to ACT, resulting in substitution of threonine for alanine (Crystal, 1990). Holmes et al. (1990) reported that the W(Bethesda) form differs from the normal M1(ala-213) allele by a change in codon 336 from GCT to ACT. Although W(Bethesda) mRNA was translated normally in vitro, transfection of the W(Bethesda) cDNA into COS-I cells was associated with AAT secretion only 50% that of cells transfected with normal cDNA. There was no intracellular accumulation as observed with the Z allele, but reduced intracellular AAT suggested degradation of newly synthesized W(Bethesda) molecules.


.0030 PI NULL(DEVON)

PI Q0(DEVON)
PI NULL(NEWPORT)
PI Q0(NEWPORT)
SERPINA1, GLY115SER
  
RCV000019597...

This variant, which is associated with increased risk of emphysema and liver disease, is due to a change in exon 2, resulting in substitution of serine for glycine-115 (Crystal, 1990). In a compound heterozygote carrying the common disease-producing mutation Pi Z (107400.0011), Graham et al. (1990) found a substitution of glycine-115 by serine. The mutation occurred on the background of M3. A change in codon 115 from GGC to AGC was responsible.


.0031 PI NULL(LUDWIGSHAFEN)

PI Q0(LUDWIGSHAFEN)
SERPINA1, ILE92ASN
  
RCV000019601...

In this variant, which is associated with increased risk of emphysema and liver disease, a change in codon 92 from ATC to AAC in exon 2 results in substitution of asparagine for isoleucine (Crystal, 1990). This substitution of a polar for a nonpolar amino acid occurs in 1 of the alpha-helices and is predicted to disrupt the tertiary structure (Fraizer et al., 1990). Fraizer et al. (1990) identified a T-to-A substitution in a German patient.


.0032 PI Z(WREXHAM)

SERPINA1, SER-19LEU
  
RCV000019570...

In a compound heterozygote with the common disease-producing PI Z mutation (107400.0011), Graham et al. (1990) found a change from TCG to TTG in codon -19, which resulted in a change from serine to leucine in the signal peptide.


.0033 MOVED TO 107400.0028


.0034 PI NULL(HONG KONG 2)

PI Q0(HONG KONG 2)
SERPINA1,
   RCV000019573

.0035 PI NULL(RIEDENBURG)

SERPINA1, DEL
   RCV000019603

Poller et al. (1991) found complete deletion of the AAT gene as the basis for PI Q0(Riedenburg). The deletion extended into the 3-prime flanking region of the gene but did not include the noncoding AAT-related gene (PIL), which is located 12 kb downstream of AAT (Hofker et al., 1988).


.0036 PI KALSHEKER-POLLER

SERPINA1, G-A, 3-PRIME UTR ENHANCER
  
RCV000019604...

Kalsheker et al. (1987) and Poller et al. (1990) reported a mutation in the 3-prime flanking sequence of the AAT gene that occurs in about 17% of patients with chronic respiratory disease. The mutation is a G-to-A nucleotide substitution in an octamer (OCT)-like sequence. Because TCGA is converted to TCAA, the mutation is detected as a restriction fragment length polymorphism with the restriction enzyme TaqI. The mutation does not appear to affect basal expression of the protein as the plasma concentration of alpha-1-antitrypsin is normal in persons who carry the mutation; however, binding and functional studies by Morgan et al. (1993) suggested that it may reduce the rise in plasma AAT concentration that occurs during inflammation. Stimulation by cytokines, such as interleukin-6 (IL6; 147620), may be lacking. Morgan et al. (1993) pointed out a precedent for such a mechanism in an unrelated gene: an enhancer element in the 3-prime flanking sequence of the erythropoietin gene increases gene expression nearly 15-fold during hypoxia.


.0037 PI P(DUARTE)

SERPINA1, ASP256VAL
   RCV000019576...

Hildesheim et al. (1993) demonstrated that the deficiency-producing change in the PI gene in P(Duarte) is the same as that in P(Lowell) (107400.0019). The alleles differ with respect to polymorphic nucleotides at other positions in the gene. They referred to this as genetic diversity from a limited repertoire of mutations on different common allelic backgrounds.


.0038 PI NULL(WEST)

PI Q0(WEST)
SERPINA1, IVS2DS, G-T, +1
  
RCV000019606...

During routine screening of individuals applying for enrollment in the US AAT Deficiency Registry, Laubach et al. (1993) identified a patient with emphysema and a PI type heterozygous for a novel AAT null allele. The novel allele, designated PI*Q0(West), was characterized by a single G-to-T transversion at position 1 of intron 2, a highly conserved nucleotide position. This resulted in an in-frame deletion of amino acids gly164-to-lys191. This was the first splicing mutation observed in the AAT gene.


.0039 PI S(IIYAMA)

SERPINA1, SER53PHE
  
RCV000019608...

In a 32-year-old Japanese male with pulmonary emphysema, Yuasa et al. (1993) demonstrated homozygosity for a C-to-T transition at codon 53 resulting in substitution of phenylalanine for serine. They commented on the fact that, in Japanese, deficiency in null alleles at the AAT locus are extremely rare and PI*Z, which occurs at polymorphic frequencies in Caucasians, has not been reported. The only other Japanese case of AAT deficiency was that due to PI M(Nichinan) (107400.0017) reported by Matsunaga et al. (1990).


.0040 PI Z(BRISTOL)

SERPINA1, THR85MET ON M1V
  
RCV000019609...

In a woman with an obstetric history of 3 perinatal deaths from fulminant liver disease and no living offspring, Lovegrove et al. (1997) found that she and her father were both heterozygotes for PI M1Z(Bristol). The Z(Bristol) protein was found to be active as a proteinase inhibitor but appeared to be deficient in the plasma to about the same degree as the S protein in MS heterozygotes. It focused on the basic side of Z and lacked the normal pattern of secondary isoforms associated with the commonly occurring AAT variants and migrated faster than normal on an SDS electrophoresis gel. The Z(Bristol) mutation was found to be a C-to-T transition at codon 85, changing ACG (thr) to ATG (met). This disrupted the N-glycosylation site starting at asn83, preventing glycosylation at residue 83 in the PI Z(Bristol) protein, and explained the protein isoelectric focusing and SDS gel electrophoresis results. An analysis of haplotypes in the propositus and her father indicated that the Z(Bristol) mutation occurred on the common M1(val213) genetic background. Of the 3 offspring with perinatal death from fulminant liver disease, 2 were by the woman's husband and 1 by an artificial insemination donor. Of the 2 offspring who were tested for the mutation, 1 had the variant and the other did not. Thus, the relationship between Z(Bristol) and fulminant liver disease in the offspring was unclear.


REFERENCES

  1. Arnaud, P., Galbraith, R. M., Galbraith, G. M. P., Allen, R. C., Fudenberg, H. H. A new allele of human alpha-1-antitrypsin: Pi (N Hampton). Am. J. Hum. Genet. 30: 653-659, 1978. [PubMed: 311584, related citations]

  2. Axelsson, U., Laurell, C. B. Hereditary variants of serum alpha-1-antitrypsin. Am. J. Hum. Genet. 17: 466-472, 1965. [PubMed: 4158556, related citations]

  3. Babron, M. C., Constans, J., Dugoujon, J. M., Cambon-Thomsen, A., Bonaiti-Pellie, C. The Gm-Pi linkage in 843 French families: effect of the alleles Pi Z and Pi S. Ann. Hum. Genet. 54: 107-113, 1990. [PubMed: 2382967, related citations] [Full Text]

  4. Bartlett, J. R., Friedman, K. J., Ling, S. C., Pace, R. G., Bell, S. C., Bourke, B., Castaldo, G., Castellani, C., Cipolli, M., Colombo, C., Colombo, J. L., Debray, D., and 22 others. Genetic modifiers of liver disease in cystic fibrosis. JAMA 302: 1076-1083, 2009. [PubMed: 19738092, related citations] [Full Text]

  5. Bender, K., Muller, C. R., Schmidt, A., Strohmaier, U., Wienker, T. F. Linkage studies on the human Pi, Gm, GLO, and HLA genes. Hum. Genet. 49: 159-166, 1979. [PubMed: 112033, related citations] [Full Text]

  6. Billingsley, G. D., Walter, M. A., Cox, D. W. Physical linkage of alpha-1-antitrypsin and alpha-1-antichymotrypsin by pulsed field gel electrophoresis. (Abstract) Am. J. Hum. Genet. 45 (suppl.): A130, 1989.

  7. Billingsley, G. D., Walter, M. A., Hammond, G. L., Cox, D. W. Physical mapping of four serpin genes: alpha-1-antitrypsin, alpha-1-antichymotrypsin, corticosteroid-binding globulin, and protein C inhibitor, within a 280-kb region on chromosome 14q32.1. Am. J. Hum. Genet. 52: 343-353, 1993. [PubMed: 8381582, related citations]

  8. Blanco, I., Bustillo, E. F., Rodriguez, M. C. Distribution of alpha-1-antitrypsin PI S and PI Z frequencies in countries outside Europe: a meta-analysis. Clin. Genet. 60: 431-441, 2001. [PubMed: 11846735, related citations] [Full Text]

  9. Boomsma, D. I., Frants, R. R., Bank, R. A., Martin, N. G. Protease inhibitor (Pi) locus, fertility and twinning. Hum. Genet. 89: 329-332, 1992. [PubMed: 1601424, related citations] [Full Text]

  10. Botstein, D., White, R. L., Skolnick, M., Davis, R. W. Construction of a genetic linkage map in man using restriction fragment length polymorphisms. Am. J. Hum. Genet. 32: 314-331, 1980. [PubMed: 6247908, related citations]

  11. Brantly, M., Courtney, M., Crystal, R. G. Repair of the secretion defect in the Z form of alpha-1-antitrypsin by addition of a second mutation. Science 242: 1700-1702, 1988. [PubMed: 2904702, related citations] [Full Text]

  12. Brennan, S. O., Carrell, R. W. Alpha-1-antitrypsin Christchurch, 363 glu-to-lys: mutation at the P-prime-5 position does not affect inhibitory activity. Biochim. Biophys. Acta 873: 13-19, 1986. [PubMed: 3527273, related citations] [Full Text]

  13. Bristow, C. L., Mercatante, D. R., Kole, R. HIV-1 preferentially binds receptors copatched with cell-surface elastase. Blood 102: 4479-4486, 2003. [PubMed: 12933574, related citations] [Full Text]

  14. Bristow, C. L., Patel, H., Arnold, R. R. Self antigen prognostic for human immunodeficiency virus disease progression. Clin. Diagn. Lab. Immun. 8: 937-942, 2001. [PubMed: 11527807, images, related citations] [Full Text]

  15. Bristow, C. L. Slow human immunodeficiency virus (HIV) infectivity correlated with low HIV coreceptor levels. Clin. Diagn. Lab. Immun. 8: 932-936, 2001. [PubMed: 11527806, images, related citations] [Full Text]

  16. Campbell, R. A., Campbell, H. D., Bircher, J. S., de Araujo, C. V., Denorme, F., Crandell, J. L., Rustad, J. L., Monts, J., Cody, M. J., Kosaka, Y., Yost, C. C. Placental HTRA1 cleaves alpha-1-antitrypsin to generate a NET-inhibitory peptide. Blood 138: 977-988, 2021. [PubMed: 34192300, images, related citations] [Full Text]

  17. Carrell, R. W., Jeppsson, J.-O., Laurell, C.-B., Brennan, S. O., Owen, M. C., Vaughan, L., Boswell, D. R. Structure and variation of human alpha-1-antitrypsin. Nature 298: 329-334, 1982. [PubMed: 7045697, related citations] [Full Text]

  18. Carrell, R. W. Alpha-1-antitrypsin: molecular pathology, leukocytes, and tissue damage. J. Clin. Invest. 78: 1427-1431, 1986. [PubMed: 3537008, related citations] [Full Text]

  19. Chappell, S., Daly, L., Morgan, K., Baranes, T. G., Roca, J., Rabinovich, R., Millar, A., Donnelly, S. C., Keatings, V., MacNee, W., Stolk, J., Hiemstra, P., Miniati, M., Monti, S., O'Connor, C. M., Kalsheker, N. Cryptic haplotypes of SERPINA1 confer susceptibility to chronic obstructive pulmonary disease. Hum. Mutat. 27: 103-109, 2006. [PubMed: 16278826, related citations] [Full Text]

  20. Chapuis-Cellier, C., Verdier, M., Lepetit, J. C., Fudenberg, H. H., Arnaud, P. Pi-Gm linkage: evidence for linkage in males but not in females and for an effect of the S allele of the Pi system. J. Immunogenet. 8: 257-262, 1981. [PubMed: 6792288, related citations] [Full Text]

  21. Clark, P., Breit, S. N., Dawkins, R. L., Penny, R. Genetic study of a family with two members with Weber-Christian disease (panniculitis) and alpha-1-antitrypsin deficiency. Am. J. Med. Genet. 13: 57-62, 1982. [PubMed: 6982619, related citations] [Full Text]

  22. Clark, P., Martin, N. G. An excess of the Pi(s) allele in dizygotic twins and their mothers. Hum. Genet. 61: 171-174, 1982. [PubMed: 6982218, related citations] [Full Text]

  23. Cox, D. W., Billingsley, G. D., Mansfield, T. DNA restriction-site polymorphisms associated with the alpha-1-antitrypsin gene. Am. J. Hum. Genet. 41: 891-906, 1987. [PubMed: 2890296, related citations]

  24. Cox, D. W., Levison, H. Emphysema of early onset associated with a complete deficiency of alpha-1-antitrypsin (null homozygotes). Am. Rev. Respir. Dis. 137: 371-375, 1988. [PubMed: 3257661, related citations] [Full Text]

  25. Cox, D. W., Markovic, V. D., Teshima, I. E. Genes for immunoglobulin heavy chains and for alpha-1-antitrypsin are localized to specific regions of chromosome 14q. Nature 297: 428-430, 1982. [PubMed: 6804874, related citations] [Full Text]

  26. Cox, D. W., Smyth, S. Risk for liver disease in adults with alpha-1-antitrypsin deficiency. Am. J. Med. 74: 221-227, 1983. [PubMed: 6600583, related citations] [Full Text]

  27. Cox, D. W., Woo, S. L. C., Mansfield, T. DNA restriction fragments associated with alpha-1-antitrypsin indicate a single origin for deficiency allele PI Z. Nature 316: 79-81, 1985. [PubMed: 2989709, related citations] [Full Text]

  28. Cox, D. W. The effect of neuraminidase on genetic variants of alpha-1-antitrypsin. Am. J. Hum. Genet. 27: 165-177, 1975. [PubMed: 1079112, related citations]

  29. Cox, D. W. Genetic variation in alpha-1-antitrypsin. (Editorial) Am. J. Hum. Genet. 30: 660-662, 1978. [PubMed: 311585, related citations]

  30. Cox, D. W. Transmission of Z allele from heterozygotes for alpha-1-antitrypsin deficiency. (Letter) Am. J. Hum. Genet. 32: 455-457, 1980. [PubMed: 17948551, related citations]

  31. Cox, D. W. New variants of alpha-1-antitrypsin: comparison of Pi typing techniques. Am. J. Hum. Genet. 33: 354-365, 1981. [PubMed: 6972696, related citations]

  32. Cox, D. W. Alpha-1-antitrypsin deficiency. In: Scriver, C. R.; Beaudet, A. L.; Sly, W. S.; Valle, D. (eds.): The Metabolic Basis of Inherited Disease. (6th ed.) New York: McGraw-Hill (pub.) 1989.

  33. Cox, D. W. Alpha-1-antitrypsin. In: Scriver, C. R.; Beaudet, A. L.; Sly, W. S.; Valle, D. (eds.): The Metabolic and Molecular Bases of Inherited Disease. Vol. IV. (8th ed.) New York: McGraw-Hill 2001. Pp. 5559-5584.

  34. Crystal, R. G. The alpha-1-antitrypsin gene and its deficiency states. Trends Genet. 5: 411-417, 1989. [PubMed: 2696185, related citations] [Full Text]

  35. Crystal, R. G. Alpha-1-antitrypsin deficiency, emphysema, and liver disease: genetic basis and strategies for therapy. J. Clin. Invest. 85: 1343-1352, 1990. [PubMed: 2185272, related citations] [Full Text]

  36. Curiel, D., Brantly, M., Curiel, E., Stier, L., Crystal, R. G. Alpha-1-antitrypsin deficiency caused by the alpha-1-antitrypsin null(Mattawa) gene: an insertion mutation rendering the alpha-1-antitrypsin gene incapable of producing alpha-1-antitrypsin. J. Clin. Invest. 83: 1144-1152, 1989. Note: Erratum: J. Clin. Invest. 84: following 1041, 1989. [PubMed: 2539391, related citations] [Full Text]

  37. Curiel, D. T., Chytil, A., Courtney, M., Crystal, R. G. Serum alpha-1-antitrypsin deficiency associated with the common S-type (glu264-to-val) mutation results from intracellular degradation of alpha-1-antitrypsin prior to secretion. J. Biol. Chem. 264: 10477-10486, 1989. [PubMed: 2567291, related citations]

  38. Curiel, D. T., Holmes, M. D., Okayama, H., Brantly, M. L., Vogelmeier, C., Travis, W. D. Stier, L. E.; Perks, W. H. and Crystal, R. G.: Molecular basis of the liver and lung disease associated with the alpha-1-antitrypsin deficiency allele M(Malton). J. Biol. Chem. 264: 13938-13945, 1989. [PubMed: 2788166, related citations]

  39. Curiel, D. T., Vogelmeier, C., Hubbard, R. C., Stier, L. E., Crystal, R. G. Molecular basis of alpha-1-antitrypsin deficiency and emphysema associated with the alpha-1-antitrypsin M(Mineral Springs) allele. Molec. Cell. Biol. 10: 47-56, 1990. [PubMed: 1967187, related citations] [Full Text]

  40. Darlington, G. J., Astrin, K. H., Muirhead, S. P., Desnick, R. J., Smith, M. Assignment of human alpha-1-antitrypsin to chromosome 14 by somatic cell hybrid analysis. Proc. Nat. Acad. Sci. 79: 870-873, 1982. [PubMed: 6801664, related citations] [Full Text]

  41. DeCroo, S., Kamboh, M. I., Ferrell, R. E. Population genetics of alpha-1-antitrypsin polymorphism in US whites, US blacks and African blacks. Hum. Hered. 41: 215-221, 1991. [PubMed: 1783408, related citations] [Full Text]

  42. Dycaico, M. J., Grant, S. G. N., Felts, K., Nichols, W. S., Geller, S. A., Hager, J. H., Pollard, A. J., Kohler, S. W., Short, H. P., Jirik, F. R., Hanahan, D., Sorge, J. A. Neonatal hepatitis induced by alpha-1-antitrypsin: a transgenic mouse model. Science 242: 1409-1415, 1988. [PubMed: 3264419, related citations] [Full Text]

  43. Emmerich, J., Alhenc-Gelas, M., Gandrille, S., Guichet, C., Fiessinger, J.-N., Aiach, M. Mechanism of protein C deficiency in a patient with arginine 358 alpha(1)-antitrypsin (Pittsburgh mutation): role in the maintenance of hemostatic balance. J. Lab. Clin. Med. 125: 531-539, 1995. [PubMed: 7706910, related citations]

  44. Faber, J. P., Poller, W., Weidinger, S., Kirchgesser, M., Schwaab, R., Bidlingmaier, S., Olek, K. Identification and DNA sequence analysis of 15 new alpha-1-antitrypsin variants, including two Pi* Q alleles and one deficient Pi* M allele. Am. J. Hum. Genet. 55: 1113-1121, 1994. [PubMed: 7977369, related citations]

  45. Faber, J. P., Weidinger, S., Goedde, H.-W., Olek, K. The deficient alpha-1-antitrypsin phenotype Pi P is associated with an A-to-T transversion in exon III of the gene. (Letter) Am. J. Hum. Genet. 45: 161-163, 1989. [PubMed: 2787118, related citations]

  46. Faber, J. P., Weidinger, S., Olek, K. Sequence data of two rare deficient alpha-1-antitrypsin phenotypes. (Abstract) Cytogenet. Cell Genet. 51: 995-996, 1989.

  47. Faber, J.-P., Weidinger, S., Olek, K. Sequence data of the rare deficient alpha-1-antitrypsin variant PI Z(Augsburg). Am. J. Hum. Genet. 46: 1158-1162, 1990. [PubMed: 2339709, related citations]

  48. Fagerhol, M. K., Braend, M. Serum prealbumin: polymorphism in man. Science 149: 986-987, 1965. [PubMed: 5827347, related citations] [Full Text]

  49. Fagerhol, M. K., Cox, D. W. The Pi polymorphism: genetic, biochemical, and clinical aspects of human alpha-1-antitrypsin. Adv. Hum. Genet. 11: 1-62, 1981. [PubMed: 6168185, related citations]

  50. Fagerhol, M. K., Gedde-Dahl, T., Jr. Genetics of the Pi serum types: family studies of the inherited variants of serum alpha-1-antitrypsin. Hum. Hered. 19: 354-359, 1969. [PubMed: 5366282, related citations] [Full Text]

  51. Fagerhol, M. K., Hauge, H. E. The Pi phenotype MP: discovery of a ninth allele belonging to the system of inherited variants of serum alpha-1-antitrypsin. Vox Sang. 15: 396-400, 1968. [PubMed: 5699691, related citations] [Full Text]

  52. Fagerhol, M. K., Laurell, C. B. The polymorphism of 'prealbumins' and alpha-1-antitrypsin in human sera. Clin. Chim. Acta 16: 199-203, 1967. [PubMed: 4166396, related citations] [Full Text]

  53. Fagerhol, M. K., Laurell, C. B. The Pi system--inherited variants of serum alpha-1-antitrypsin. Prog. Med. Genet. 7: 96-111, 1970. [PubMed: 4911922, related citations]

  54. Fagerhol, M. K., Tenfjord, O. W. Serum Pi types in some European, American, Asian and African populations. Acta Path. Microbiol. Scand. 72: 601-608, 1968. [PubMed: 5681806, related citations] [Full Text]

  55. Fagerhol, M. K. The Pi system: genetic variants of serum alpha-1-antitrypsin. Ser. Haemat. 1: 153-161, 1968.

  56. Fraizer, G. C., Harrold, T. R., Hofker, M. H., Cox, D. W. In-frame single codon deletion in the M(Malton) deficiency allele of alpha-1-antitrypsin. Am. J. Hum. Genet. 44: 894-902, 1989. [PubMed: 2786335, related citations]

  57. Fraizer, G. C., Siewertsen, M. A., Hofker, M. H., Brubacher, M. G., Cox, D. W. A null deficiency allele of alpha-1-antitrypsin, Q0-Ludwigshafen, with altered tertiary structure. J. Clin. Invest. 86: 1878-1884, 1990. [PubMed: 2254451, related citations] [Full Text]

  58. Fraizer, G. C., Siewertsen, M., Harold, T. R., Cox, D. W. Deletion/frameshift mutation in the alpha-1-antitrypsin null allele, PI*Q0(Bolton). Hum. Genet. 83: 377-382, 1989. [PubMed: 2807278, related citations] [Full Text]

  59. Frants, R., Eriksson, A. W. A new unstable Pi M variant of alpha-1-antitrypsin in a Finnish isolate. Hum. Hered. 30: 333-342, 1980. [PubMed: 6971248, related citations] [Full Text]

  60. Garver, R. I., Jr., Mornex, J.-F., Nukiwa, T., Brantly, M., Courtney, M., LeCocq, J.-P., Crystal, R. G. Alpha-1-antitrypsin deficiency and emphysema caused by homozygous inheritance of non-expressing alpha-1-antitrypsin genes. New Eng. J. Med. 314: 762-766, 1986. [PubMed: 3485249, related citations] [Full Text]

  61. Gedde-Dahl, T., Jr., Cook, P. J. L., Fagerhol, M. K., Pierce, J. A. Improved estimate of the Gm-Pi linkage. Ann. Hum. Genet. 39: 43-50, 1975. [PubMed: 810069, related citations] [Full Text]

  62. Gedde-Dahl, T., Jr., Cook, P. J. L., Fagerhol, M. K., Pierce, J. A. The Gm-Pi linkage: a summary estimate. Birth Defects Orig. Art. Ser. XI(3): 157-158, 1975. Note: Alternate: Cytogenet. Cell Genet. 14: 327-328, 1975.

  63. Gedde-Dahl, T., Jr., Fagerhol, M. K., Cook, P. J. L., Noades, J. Autosomal linkage between the Gm and Pi loci in man. Ann. Hum. Genet. 35: 393-400, 1972. [PubMed: 5073686, related citations]

  64. Gedde-Dahl, T., Jr., Frants, R., Olaisen, B., Eriksson, A. W., van Loghem, E., Lamm, L. The Gm-Pi linkage heterogeneity in view of Pi M subtypes. Ann. Hum. Genet. 45: 143-153, 1981. [PubMed: 6797346, related citations] [Full Text]

  65. Graham, A., Hayes, K., Weidinger, S., Newton, C. R., Markham, A. F., Kalsheker, N. A. Characterisation of the alpha-1-antitrypsin M3 gene, a normal variant. Hum. Genet. 85: 381-382, 1990. [PubMed: 2394452, related citations] [Full Text]

  66. Graham, A., Kalsheker, N. A., Bamforth, F. J., Newton, C. R., Markham, A. F. Molecular characterization of two alpha-1-antitrypsin deficiency variants: proteinase inhibitor (Pi) null(Newport) (gly(115)-to-ser) and (Pi) Z Wrexham (ser(-19)-to-leu). Hum. Genet. 85: 537-540, 1990. [PubMed: 2227940, related citations] [Full Text]

  67. Graham, A., Kalsheker, N. A., Newton, C. R., Bamforth, F. J., Powell, S. J., Markham, A. F. Molecular characterisation of three alpha-1-antitrypsin deficiency variants: proteinase inhibitor (Pi) Null (Cardiff) (asp256-to-val), Pi M(Malton) (phe51-deletion) and Pi I (arg39-to-cys). Hum. Genet. 84: 55-58, 1989. [PubMed: 2606478, related citations] [Full Text]

  68. Green, C., Brown, G., Dafforn, T. R., Reichhart, J.-M., Morley, T., Lomas, D. A., Gubb, D. Drosophila necrotic mutations mirror disease-associated variants of human serpins. Development 130: 1473-1478, 2003. [PubMed: 12588861, related citations] [Full Text]

  69. Hafeez, W., Ciliberto, G., Perlmutter, D. H. Constitutive and modulated expression of the human alpha-1 antitrypsin gene: different transcriptional initiation sites used in three different cell types. J. Clin. Invest. 89: 1214-1222, 1992. [PubMed: 1556183, related citations] [Full Text]

  70. Hidvegi, T., Ewing, M., Hale, P., Dippold, C., Beckett, C., Kemp, C., Maurice, N., Mukherjee, A., Goldbach, C., Watkins, S., Michalopoulos, G., Perlmutter, D. H. An autophagy-enhancing drug promotes degradation of mutant alpha-1-antitrypsin Z and reduces hepatic fibrosis. Science 329: 229-232, 2010. [PubMed: 20522742, related citations] [Full Text]

  71. Hildesheim, J., Kinsley, G., Bissell, M., Pierce, J., Brantly, M. Genetic diversity from a limited repertoire of mutations on different common allelic backgrounds: alpha-1-antitrypsin deficiency variant P(Duarte). Hum. Mutat. 2: 221-228, 1993. [PubMed: 8364590, related citations] [Full Text]

  72. Hofker, M. H., Nelen, M., Klasen, E. C., Nukiwa, T., Curiel, D., Crystal, R. G., Frants, R. R. Cloning and characterization of an alpha-1-antitrypsin like gene 12 kb downstream of the genuine alpha-1-antitrypsin gene. Biochem. Biophys. Res. Commun. 155: 634-642, 1988. [PubMed: 2901833, related citations] [Full Text]

  73. Hofker, M. H., Nukiwa, T., van Paassen, H. M. B., Nelen, M., Kramps, J. A., Klasen, E. C., Frants, R. R., Crystal, R. G. A pro-to-leu substitution in codon 369 of the alpha-1-antitrypsin deficiency variant PI MHeerlen. Hum. Genet. 81: 264-268, 1989. [PubMed: 2784123, related citations] [Full Text]

  74. Holmes, M. D., Brantly, M. L., Crystal, R. G. Molecular analysis of the heterogeneity among the P-family of alpha-1-antitrypsin alleles. Am. Rev. Resp. Dis. 142: 1185-1192, 1990. [PubMed: 2240842, related citations] [Full Text]

  75. Holmes, M. D., Brantly, M. L., Curiel, D. T., Weidinger, S., Crystal, R. G. Characterization of the normal alpha-1-antitrypsin allele V(Munich): a variant associated with a unique protein isoelectric focusing pattern. Am. J. Hum. Genet. 46: 810-816, 1990. [PubMed: 2316526, related citations]

  76. Holmes, M. D., Brantly, M. L., Fells, G. A., Crystal, R. G. Alpha-1-antitrypsin W(Bethesda): molecular basis of an unusual alpha-1-antitrypsin deficiency variant. Biochem. Biophys. Res. Commun. 170: 1013-1020, 1990. [PubMed: 2390072, related citations] [Full Text]

  77. Hug, G., Chuck, G., Fagerhol, M. K. Pi(P-Clifton): a new alpha(1) antitrypsin allele in an American Negro family. J. Med. Genet. 18: 43-45, 1981. [PubMed: 6973024, related citations] [Full Text]

  78. Hug, G., Chuck, G., Slemmer, T. M., Fagerhol, M. K. Pi (Ecincinnati): a new alpha-1-antitrypsin allele in three Negro families. Hum. Genet. 54: 361-364, 1980. [PubMed: 6967446, related citations] [Full Text]

  79. Hutchison, D. C. S. Alpha-1-antitrypsin deficiency in Europe: geographical distribution of Pi types S and Z. Respir. Med. 92: 367-377, 1998. [PubMed: 9692092, related citations] [Full Text]

  80. Iammarino, R. M., Wagener, D. K., Allen, R. C. Segregation distortion of the alpha-1-antitrypsin Pi Z allele. Am. J. Hum. Genet. 31: 508-517, 1979. [PubMed: 314754, related citations]

  81. Kalsheker, N. A., Hodgson, I. J., Watkins, G. L., White, J. P., Morrison, H. M., Stockley, R. A. Deoxyribonucleic acid (DNA) polymorphism of the alpha-1-antitrypsin gene in chronic lung disease. Brit. Med. J. 294: 1511-1514, 1987. [PubMed: 3038256, related citations] [Full Text]

  82. Kalsheker, N., Hayes, K., Weidinger, S., Graham, A. What is Pi (proteinase inhibitor) null or PiQ0? A problem highlighted by the alpha-1-antitrypsin M(Heerlen) mutation. J. Med. Genet. 29: 27-29, 1992. [PubMed: 1552539, related citations] [Full Text]

  83. Kamimoto, T., Shoji, S., Hidvegi, T., Mizushima, N., Umebayashi, K., Perlmutter, D. H., Yoshimori, T. Intracellular inclusions containing mutant alpha(1)-antitrypsin Z are propagated in the absence of autophagic activity. J. Biol. Chem. 281: 4467-4476, 2006. [PubMed: 16365039, related citations] [Full Text]

  84. Kawakami, Y., Irie, T., Kishi, F., Asanuma, Y., Shida, A., Yoshikawa, T., Kamishima, K., Hasegawa, H., Murao, M. Familial aggregation of abnormal ventilatory control and pulmonary function in chronic obstructive pulmonary disease. Europ. J. Resp. Dis. 62: 56-64, 1981. [PubMed: 7227484, related citations]

  85. Kelly, E., Greene, C. M., Carroll, T. P., McElvaney, N. G., O'Neill, S. J. Selenoprotein S/SEPS1 modifies endoplasmic reticulum stress in Z variant alpha-1-antitrypsin deficiency. J. Biol. Chem. 284: 16891-16897, 2009. [PubMed: 19398551, images, related citations] [Full Text]

  86. Keyeux, G., Gilgenkrantz, S., Lefranc, G., Lefranc, M.-P. Molecular characterization of a ring chromosome 14 showing that the PI locus is centromeric to the D14S1 and IGH loci. Hum. Genet. 82: 219-222, 1989. [PubMed: 2543620, related citations] [Full Text]

  87. Kramps, J. A., Brouwers, J. W., Maesen, F., Dijkman, J. H. Pi(M-Heerlen), a Pi(M) allele resulting in very low alpha-1-antitrypsin serum levels. Hum. Genet. 59: 104-107, 1981. [PubMed: 6976926, related citations] [Full Text]

  88. Kueppers, F., Bearn, A. G. An inherited alpha-1-antitrypsin variant. Humangenetik 4: 217-220, 1967. [PubMed: 6080804, related citations] [Full Text]

  89. Kueppers, F., Christopherson, M. J. Alpha-1-antitrypsin: further genetic heterogeneity revealed by isoelectric focusing. Am. J. Hum. Genet. 30: 359-365, 1978. [PubMed: 309724, related citations]

  90. Kurachi, K., Chandra, T., Degen, S. J. F., White, T. T., Marchioro, T. L., Woo, S. L. C., Davie, E. W. Cloning and sequence of cDNA coding for alpha-1-antitrypsin. Proc. Nat. Acad. Sci. 78: 6826-6830, 1981. [PubMed: 7031661, related citations] [Full Text]

  91. Lace, B., Sveger, T., Krams, A., Cernevska, G., Krumina, A. Age of SERPINA1 gene PI Z mutation: Swedish and Latvian population analysis. Ann. Hum. Genet. 72: 300-304, 2008. [PubMed: 18294358, related citations] [Full Text]

  92. Lai, E. C., Kao, F.-T., Law, M. L., Woo, S. L. C. Assignment of the alpha-1-antitrypsin gene and a sequence-related gene to human chromosome 14 by molecular hybridization. Am. J. Hum. Genet. 35: 385-392, 1983. [PubMed: 6602546, related citations]

  93. Laubach, V. E., Ryan, W. J., Brantly, M. Characterization of a human alpha-1-antitrypsin null allele involving aberrant mRNA splicing. Hum. Molec. Genet. 2: 1001-1005, 1993. [PubMed: 8364536, related citations] [Full Text]

  94. Ledbetter, S. A., Ledbetter, D. H., Ledley, F. D., Woo, S. Localization of phenylalanine hydroxylase (PAH) and alpha-1 antitrypsin (AAT) loci in mouse genome by synteny and in situ hybridization. (Abstract) Am. J. Hum. Genet. 41: A173, 1987.

  95. Lewis, J. H., Iammarino, R. M., Spero, J. A., Hasiba, U. Antithrombin Pittsburgh: an alpha-1-antitrypsin variant causing hemorrhagic disease. Blood 51: 129-137, 1978. [PubMed: 412531, related citations]

  96. Lieberman, J., Borhani, N. O., Feinleib, M. Alpha-1-antitrypsin deficiency in twins and parents-of-twins. Clin. Genet. 15: 29-36, 1979. [PubMed: 310370, related citations] [Full Text]

  97. Long, G. L., Chandra, T., Woo, S. L. C., Davie, E. W., Kurachi, K. Complete sequence of the cDNA for human alpha-1-antitrypsin and the gene for the S variant. Biochemistry 23: 4828-4837, 1984. [PubMed: 6093867, related citations] [Full Text]

  98. Lopez, V., Oetliker, O., Colombo, J. P., Butler, R. Ein Fall von familiaerem alpha-1-Antitrypsinmangel. Helv. Paediat. Acta 19: 296-303, 1964. [PubMed: 14229909, related citations]

  99. Lovegrove, J. U., Jeremiah, S., Gillett, G. T., Temple, I. K., Povey, S., Whitehouse, D. B. A new alpha 1-antitrypsin mutation, thr-met 85, (PI Z-Bristol) associated with novel electrophoresis properties. Ann. Hum. Genet. 61: 385-391, 1997. [PubMed: 9459000, related citations] [Full Text]

  100. Martorana, P. A., Brand, T., Gardi, C., van Even, P., de Santi, M. M., Calzoni, P., Marcolongo, P., Lungarella, G. The pallid mouse: a model of genetic alpha-1-antitrypsin deficiency. Lab. Invest. 68: 233-241, 1993. [PubMed: 8441253, related citations]

  101. Matsunaga, E., Shiokawa, S., Nakamura, H., Maruyama, T., Tsuda, K., Fukumaki, Y. Molecular analysis of the gene of the alpha-1-antitrypsin deficiency variant, M(Nichinan). Am. J. Hum. Genet. 46: 602-612, 1990. [PubMed: 2309708, related citations]

  102. Morgan, K., Scobie, G., Kalsheker, N. A. Point mutation in a 3-prime flanking sequence of the alpha-1-antitrypsin gene associated with chronic respiratory disease occurs in a regulatory sequence. Hum. Molec. Genet. 2: 253-257, 1993. [PubMed: 8499914, related citations] [Full Text]

  103. Munch, J., Standker, L., Adermann, K., Schulz, A., Schindler, M., Chinnadurai, R., Pohlmann, S., Chaipan, C., Biet, T., Peters, T., Meyer, B., Wilhelm, D., Lu, H., Jing, W., Jiang, S., Forssmann, W.-G., Kirchhoff, F. Discovery and optimization of a natural HIV-1 entry inhibitor targeting the gp41 fusion peptide. Cell 129: 263-275, 2007. [PubMed: 17448989, related citations] [Full Text]

  104. Nakamura, H., Ogawa, A., Hisano, S., Fukuma, M., Tachibana, N., Tsuda, K. A family with a new deficient variant of alpha-1-antitrypsin PiM(Nichinan): with special reference to diastase-resistant, periodic acid-Schiff positive globules in the liver cells. J. Jpn. Soc. Intern. Med. 69: 47-54, 1980.

  105. Nukiwa, T., Brantly, M., Garver, R., Paul, L., Courtney, M., LeCocq, J.-P., Crystal, R. G. Evaluation of 'at risk' alpha-1-antitrypsin genotype SZ with synthetic oligonucleotide gene probes. J. Clin. Invest. 77: 528-537, 1986. [PubMed: 3484754, related citations] [Full Text]

  106. Nukiwa, T., Brantly, M. L., Ogushi, F., Fells, G. A., Crystal, R. G. Characterization of the gene and protein of the common alpha-1-antitrypsin normal M2 allele. Am. J. Hum. Genet. 43: 322-330, 1988. [PubMed: 2901226, related citations]

  107. Nukiwa, T., Satoh, K., Brantly, M. L., Ogushi, F., Fells, G. A., Courtney, M., Crystal, R. G. Identification of a second mutation in the protein-coding sequence of the Z type alpha-1-antitrypsin gene. J. Biol. Chem. 261: 15989-15994, 1986. Note: Erratum: J. Biol. Chem. 262: 10412 only, 1987. [PubMed: 3491072, related citations]

  108. Nukiwa, T., Takahashi, H., Brantly, M., Courtney, M., Crystal, R. G. Alpha-1-antitrypsin null (Granite Falls), a nonexpressing alpha-1-antitrypsin gene associated with a frameshift to stop mutation in a coding exon. J. Biol. Chem. 262: 11999-12004, 1987. [PubMed: 3040726, related citations]

  109. Okayama, H., Brantly, M., Holmes, M., Crystal, R. G. Characterization of the molecular basis of the alpha-1-antitrypsin F allele. Am. J. Hum. Genet. 48: 1154-1158, 1991. [PubMed: 2035534, related citations]

  110. Owen, M. C., Brennan, S. O., Lewis, J. H., Carrell, R. W. Mutation of antitrypsin to antithrombin: alpha-1-antitrypsin Pittsburgh (358 met-to-arg), a fatal bleeding disorder. New Eng. J. Med. 309: 694-698, 1983. [PubMed: 6604220, related citations] [Full Text]

  111. Owen, M. C., Carrell, R. W., Brennan, S. O. The abnormality of the S variant of human alpha-1-antitrypsin. Biochim. Biophys. Acta 453: 257-261, 1976. [PubMed: 1087161, related citations] [Full Text]

  112. Owen, M. C., Carrell, R. W. Alpha-1-antitrypsin: molecular abnormality of S variant. Brit. Med. J. 1: 130-131, 1976. [PubMed: 1082356, related citations] [Full Text]

  113. Pearson, S., Tetri, P., George, D. L., Francke, U. Alpha-1-antitrypsin (PI) expression in rat hepatoma-human somatic cell hybrids: evidence for PI locus on chromosome 14 and for regulatory locus on the X chromosome. (Abstract) Am. J. Hum. Genet. 33: 148A, 1981.

  114. Perlino, E., Cortese, R., Ciliberto, G. The human alpha-1-antitrypsin gene is transcribed from two different promoters in macrophages and hepatocytes. EMBO J. 6: 2767-2771, 1987. [PubMed: 3500042, related citations] [Full Text]

  115. Poller, W., Faber, J.-P., Weidinger, S., Olek, K. DNA polymorphisms associated with a new alpha-1-antitrypsin PI Q0 variant (PI Q0-Riedenburg). Hum. Genet. 86: 522-524, 1991. [PubMed: 1673114, related citations] [Full Text]

  116. Poller, W., Meisen, C., Olek, K. DNA polymorphisms of the alpha-1-antitrypsin gene region in patients with chronic obstructive pulmonary disease. Europ. J. Clin. Invest. 20: 1-7, 1990. [PubMed: 1969347, related citations] [Full Text]

  117. Roychoudhury, A. K., Nei, M. Human Polymorphic Genes: World Distribution. New York: Oxford Univ. Press (pub.) 1988. Pp. 132-135.

  118. Satoh, K., Nukiwa, T., Brantly, M., Garver, R. I., Jr., Hofker, M., Courtney, M., Crystal, R. G. Emphysema associated with complete absence of alpha-1-antitrypsin in serum and the homozygous inheritance of a stop codon in an alpha-1-antitrypsin-coding exon. Am. J. Hum. Genet. 42: 77-83, 1988. Note: Erratum: Am. J. Hum. Genet. 42: 789 only, 1988. [PubMed: 3257351, related citations]

  119. Schapira, M., Ramus, M.-A., Jallat, S., Carvallo, D., Courtney, M. Recombinant alpha-1-antitrypsin Pittsburgh (met-358 to arg) is a potent inhibitor of plasma kallikrein and activated factor XII fragment. J. Clin. Invest. 77: 635-637, 1986. [PubMed: 3484756, related citations] [Full Text]

  120. Schroeder, W. T., Miller, M. F., Woo, S. L. C., Saunders, G. F. Chromosomal localization of the human alpha-antitrypsin gene (PI) to 14q31-32. Am. J. Hum. Genet. 37: 868-872, 1985. [PubMed: 3876766, related citations]

  121. Scott, C. F., Carrell, R. W., Glaser, C. B., Kueppers, F., Lewis, J. H., Colman, R. W. Alpha-1-antitrypsin-Pittsburgh: a potent inhibitor of human plasma factor XIa, kallikrein, and factor XII. J. Clin. Invest. 77: 631-634, 1986. [PubMed: 3484755, related citations] [Full Text]

  122. Sefton, L., Kearney, P., Kelsey, G., Povey, S., Wolfe, J. Physical linkage of the genes PI and AACT. (Abstract) Cytogenet. Cell Genet. 51: 1076, 1989.

  123. Seixas, S., Mendonca, C., Costa, F., Rocha, J. Alpha-1-antitrypsin null alleles: evidence for the recurrence of the L353fsX376 mutation and a novel G-to-A transition in position +1 of intron IC affecting normal mRNA splicing. Clin. Genet. 62: 175-180, 2002. [PubMed: 12220457, related citations] [Full Text]

  124. Seri, M., Magi, B., Cellesi, C., Olia, P. M., Renieri, A., De Marchi, M. Molecular characterization of the P and I variants of alpha-1-antitrypsin. Int. J. Clin. Lab. Res. 22: 119-121, 1992. [PubMed: 1504305, related citations] [Full Text]

  125. Shapiro, L., Pott, G. B., Ralston, A. H. Alpha-1-antitrypsin inhibits human immunodeficiency virus type 1. FASEB J. 15: 115-122, 2001. [PubMed: 11149899, related citations] [Full Text]

  126. Sifers, R. N., Brashears-Macatee, S., Kidd, V. J., Muensch, H., Woo, S. L. C. A frameshift mutation results in a truncated alpha-1-antitrypsin that is retained within the rough endoplasmic reticulum. J. Biol. Chem. 263: 7330-7335, 1988. [PubMed: 3259232, related citations]

  127. Soutoglou, E., Talianidis, I. Coordination of PIC assembly and chromatin remodeling during differentiation-induced gene activation. Science 295: 1901-1904, 2002. [PubMed: 11884757, related citations] [Full Text]

  128. Takahashi, H., Crystal, R. G. Alpha-1-antitrypsin Null(Isola di Procida): an alpha-1-antitrypsin deficiency allele caused by deletion of all alpha-1-antitrypsin coding exons. Am. J. Hum. Genet. 47: 403-413, 1990. [PubMed: 1975477, related citations]

  129. Takahashi, H., Nukiwa, T., Satoh, K., Ogushi, F., Brantly, M., Fells, G., Stier, L., Courtney, M., Crystal, R. G. Characterization of the gene and protein of the alpha-1-antitrypsin 'deficiency' allele M(procida). J. Biol. Chem. 263: 15528-15534, 1988. [PubMed: 3262617, related citations]

  130. Turleau, C., de Grouchy, J., Chavin-Colin, F., Dore, F., Seger, J., Dautzenberg, M., Arthuis, M., Jeanson, C. Two patients with interstitial del(14q), one with features of Holt-Oram syndrome: exclusion mapping of PI (alpha-1-antitrypsin). Ann. Genet. 27: 237-240, 1984. [PubMed: 6335371, related citations]

  131. van Pel, M., van Os, R., Velders, G. A., Hagoort, H., Heegaard, P. M. H., Lindley, I. J. D., Willemze, R., Fibbe, W. E. Serpina1 is a potent inhibitor of IL-8-induced hematopoietic stem cell mobilization. Proc. Nat. Acad. Sci. 103: 1469-1474, 2006. [PubMed: 16432201, images, related citations] [Full Text]

  132. Vidaud, D., Emmerich, J., Alhenc-Gelas, M., Yvart, J., Fiessinger, J. N., Aiach, M. Met358-to-arg mutation of alpha-1-antitrypsin associated with protein C deficiency in a patient with mild bleeding tendency. J. Clin. Invest. 89: 1537-1543, 1992. [PubMed: 1569192, related citations] [Full Text]

  133. Weidinger, S., Jahn, W., Cujnik, F., Schwarzfischer, F. Alpha-1-antitrypsin: evidence for a fifth PI M subtype and a new deficiency allele PI*Z(Augsburg). Hum. Genet. 71: 27-29, 1985. [PubMed: 3875547, related citations] [Full Text]

  134. Weitkamp, L. R., Cox, D., Guttormsen, S., Johnston, E., Hempfling, S. Allelic specific heterogeneity in the Pi-Gm linkage group. Cytogenet. Cell Genet. 22: 647-650, 1978. [PubMed: 313312, related citations] [Full Text]

  135. Welch, S. G., McGregor, I. A., Williams, K. Alpha-1-antitrypsin (Pi) phenotypes in a village population from the Gambia, West Africa. Hum. Genet. 53: 233-235, 1980. [PubMed: 6965658, related citations] [Full Text]

  136. Whitehouse, D. B., Abbott, C. M., Lovegrove, J. U., McIntosh, I., McMahon, C. J., Mieli-Vergani, G., Mowat, A. P., Hopkinson, D. A. Genetic studies on a new deficiency gene (PI*Z-Tun) at the PI locus. J. Med. Genet. 26: 744-749, 1989. [PubMed: 2575668, related citations] [Full Text]

  137. Wilkie, A. O. M. The molecular basis of genetic dominance. J. Med. Genet. 31: 89-98, 1994. [PubMed: 8182727, related citations] [Full Text]

  138. Yamamoto, Y., Sawa, R., Okamoto, N., Matsui, A., Yanagisawa, M., Ikemoto, S. Deletion 14q(q24.3 to q32.1) syndrome: significance of peculiar facial appearance in its diagnosis, and deletion mapping of Pi (alpha-1-antitrypsin). Hum. Genet. 74: 190-192, 1986. [PubMed: 3490426, related citations] [Full Text]

  139. Yoshida, A., Chillar, R., Taylor, J. C. An alpha-1-antitrypsin variant, PiB Alhambra (lys-to-asp, glu-to-asp), with rapid anodal electrophoretic mobility. Am. J. Hum. Genet. 31: 555-563, 1979. [PubMed: 315708, related citations]

  140. Yoshida, A., Ewing, C., Wessels, M., Lieberman, J., Gaidulis, L. Molecular abnormality of Pi S variant of human alpha-1-antitrypsin. Am. J. Hum. Genet. 29: 233-239, 1977. [PubMed: 301355, related citations]

  141. Yoshida, A., Lieberman, J., Gaidulis, L., Ewing, C. Molecular abnormality of human alpha-1-antitrypsin variant (Pi-ZZ) associated with plasma activity deficiency. Proc. Nat. Acad. Sci. 73: 1324-1328, 1976. [PubMed: 1083527, related citations] [Full Text]

  142. Yoshida, A., Taylor, J. C., Van den Brock, W. G. M. Structural difference between the normal PiM(1) and the common PiM(2) variant of human alpha-1-antitrypsin. Am. J. Hum. Genet. 31: 564-568, 1979. [PubMed: 315709, related citations]

  143. Yuasa, I., Sugimoto, Y., Ichinose, M., Matsumoto, Y., Fukumaki, Y., Sasaki, T., Okada, K. PI*S(Iiyama), a deficiency gene of alpha-1-antitrypsin: evidence for the occurrence in western Japan. Jpn. J. Hum. Genet. 38: 185-191, 1993. [PubMed: 8358043, related citations] [Full Text]


Bao Lige - updated : 08/31/2022
Patricia A. Hartz - updated : 7/11/2012
Ada Hamosh - updated : 9/7/2011
Ada Hamosh - updated : 8/17/2010
Cassandra L. Kniffin - updated : 1/13/2009
Paul J. Converse - updated : 8/17/2007
Paul J. Converse - updated : 3/24/2006
Victor A. McKusick - updated : 1/20/2006
Paul J. Converse - updated : 3/14/2005
Victor A. McKusick - updated : 12/4/2003
Victor A. McKusick - updated : 7/10/2003
Ada Hamosh - updated : 4/15/2003
Victor A. McKusick - updated : 11/26/2002
Victor A. McKusick - updated : 6/3/2002
Ada Hamosh - updated : 4/2/2002
Victor A. McKusick - updated : 2/12/2002
Victor A. McKusick - updated : 1/10/2002
Victor A. McKusick - updated : 10/21/1999
Victor A. McKusick - updated : 12/10/1998
Victor A. McKusick - updated : 11/5/1998
Victor A. McKusick - updated : 7/10/1998
Cynthia K. Ewing - updated : 10/23/1996
Creation Date:
Victor A. McKusick : 6/4/1986
carol : 09/30/2022
mgross : 08/31/2022
carol : 07/20/2016
carol : 07/18/2016
carol : 7/15/2016
carol : 7/14/2016
carol : 7/9/2016
carol : 7/9/2016
alopez : 8/7/2013
carol : 5/21/2013
terry : 4/4/2013
terry : 3/15/2013
carol : 8/15/2012
terry : 8/8/2012
terry : 7/11/2012
carol : 5/30/2012
terry : 10/10/2011
mgross : 10/3/2011
alopez : 9/8/2011
terry : 9/7/2011
terry : 9/7/2010
terry : 9/7/2010
terry : 9/7/2010
alopez : 8/18/2010
joanna : 8/17/2010
terry : 8/17/2010
carol : 8/13/2010
carol : 9/15/2009
terry : 6/3/2009
wwang : 4/1/2009
wwang : 1/16/2009
ckniffin : 1/13/2009
terry : 1/8/2009
terry : 1/8/2009
terry : 1/7/2009
terry : 1/7/2009
carol : 10/4/2007
mgross : 8/17/2007
mgross : 3/29/2006
terry : 3/24/2006
carol : 2/14/2006
alopez : 1/25/2006
terry : 1/20/2006
carol : 11/28/2005
mgross : 3/14/2005
mgross : 3/14/2005
mgross : 3/17/2004
alopez : 12/10/2003
terry : 12/4/2003
carol : 11/24/2003
carol : 7/21/2003
terry : 7/10/2003
alopez : 4/17/2003
terry : 4/15/2003
cwells : 11/26/2002
terry : 11/20/2002
cwells : 6/26/2002
terry : 6/3/2002
carol : 4/25/2002
alopez : 4/5/2002
terry : 4/2/2002
terry : 3/13/2002
terry : 2/12/2002
cwells : 1/25/2002
cwells : 1/16/2002
terry : 1/10/2002
alopez : 11/23/1999
carol : 10/21/1999
carol : 12/16/1998
terry : 12/10/1998
carol : 11/16/1998
terry : 11/5/1998
dkim : 7/24/1998
dkim : 7/24/1998
carol : 7/15/1998
terry : 7/10/1998
terry : 5/29/1998
terry : 1/29/1998
mark : 10/14/1997
mark : 10/14/1997
dholmes : 8/14/1997
alopez : 7/31/1997
jenny : 7/9/1997
joanna : 7/3/1997
joanna : 6/24/1997
terry : 1/10/1997
mark : 12/29/1996
jamie : 10/23/1996
jamie : 10/16/1996
jamie : 10/14/1996
mark : 3/25/1996
terry : 7/10/1995
mark : 6/13/1995
pfoster : 5/2/1995
davew : 8/18/1994
jason : 6/17/1994
warfield : 4/4/1994

* 107400

SERPIN PEPTIDASE INHIBITOR, CLADE A, MEMBER 1; SERPINA1


Alternative titles; symbols

ALPHA-1-ANTITRYPSIN; AAT
PROTEASE INHIBITOR 1; PI
PI1
ANTI-ELASTASE
ANTITRYPSIN


HGNC Approved Gene Symbol: SERPINA1

Cytogenetic location: 14q32.13     Genomic coordinates (GRCh38): 14:94,376,747-94,390,635 (from NCBI)


Gene-Phenotype Relationships

Location Phenotype Phenotype
MIM number
Inheritance Phenotype
mapping key
14q32.13 Emphysema due to AAT deficiency 613490 Autosomal recessive 3
Emphysema-cirrhosis, due to AAT deficiency 613490 Autosomal recessive 3
Hemorrhagic diathesis due to antithrombin Pittsburgh 613490 Autosomal recessive 3

TEXT

Description

The SERPINA1 gene encodes alpha-1-antitrypsin (AAT), also known as protease inhibitor (PI), a major plasma serine protease inhibitor. AAT complexes predominantly with elastase, but also with trypsin, chymotrypsin, thrombin, and bacterial proteases. The most important inhibitory action of AAT is that against neutrophil elastase (ELANE, or HLE; 130130), a protease that degrades elastin of the alveolar walls as well as other structural proteins of a variety of tissues (review by Cox, 2001).


Cloning and Expression

Kurachi et al. (1981) cloned a nearly full-length baboon AAT cDNA (approximately 1,352 bp) and a partial human AAT cDNA (approximately 306 bp). They found more than 96% homology between the cDNA and predicted amino acid sequences of AAT in the 2 species. Comparison of baboon AAT, human antithrombin III (107300), and chicken ovalbumin indicated about 30% homology of amino acid sequence.

Long et al. (1984) cloned a full-length human AAT cDNA from a liver cDNA library. Sequence analysis revealed a precursor molecule containing a 24-amino acid signal peptide and a mature protein of 394 amino acids. AAT is primarily synthesized in the liver.

Crystal (1990) noted that hepatocytes are the major source of AAT, but that the gene is also expressed in mononuclear phagocytes and neutrophils.


Gene Structure

Lai et al. (1983) showed that the AAT gene contains 3 introns in the peptide-coding region.

Long et al. (1984) found that the genomic length of the PI gene is 10.2 kb with a 1,434-bp coding region. The gene has 4 introns; exon 1, the 5-prime portion of exon 2, and the 3-prime portion of exon 5 are noncoding regions. The first intron, 5.3 kb long, contains a 143-amino acid open reading frame (which does not appear to be an actual protein coding region), an Alu family sequence, and a pseudotranscription initiation region.

Perlino et al. (1987) found that the AAT gene in macrophages is transcribed from a macrophage-specific promoter located about 2,000 bp upstream of the hepatocyte-specific promoter. Transcription from the 2 AAT promoters is mutually exclusive; the macrophage promoter is silent in hepatocytes and the hepatocyte promoter is silent in macrophages. In macrophages, 2 distinct mRNAs are generated by alternative splicing.

Hafeez et al. (1992) demonstrated that the AAT gene has 3 macrophage-specific transcriptional initiation sites upstream from a single hepatocyte-specific transcriptional initiation site. Macrophages use these sites during basal and modulated expression. Hepatoma cells use the hepatocyte-specific transcriptional initiation site during basal and modulated expression but also switch on transcription from the upstream macrophage transcriptional initiation sites during modulation by the acute phase mediator interleukin-6 (IL6; 147620).

Soutoglou and Talianidis (2002) analyzed the ordered recruitment of factors to the human alpha-1-antitrypsin promoter around the initial activation of the gene during enterocyte differentiation. They found that a complete preinitiation complex, including phosphorylated RNA Pol II (180660), was assembled at the promoter long before transcriptional activation. The histone acetyltransferases CBP (600140) and P/CAF (602303) were recruited subsequently, but local histone hyperacetylation was delayed. After transient recruitment of the human Brahma homolog BRM (600014), remodeling of the neighboring nucleosome coincided with transcription initiation. Soutoglou and Talianidis (2002) concluded that, at this promoter, chromatin reconfiguration is a defining step of the initiation process, acting after the assembly of the Pol II machinery.


Mapping

Lai et al. (1983) used a cloned AAT gene as a hybridization probe to analyze EcoRI-digested genomic DNA from different individuals and identified 2 distinct bands (9.6 kb and 8.5 kb long) in every case. Analysis using only intronic DNA as probe showed that the authentic gene resides in the 9.6-kb fragment. The 8.5-kb fragment was thought to contain a gene with close sequence homology to that of AAT.

By studying hybrids of mouse or rat hepatoma cells with human lymphocytes, Darlington et al. (1982) and Pearson et al. (1981) achieved direct assignment of the PI locus to chromosome 14. From study of 2 families with abnormalities of the long arm of chromosome 14, Cox et al. (1982) localized GM to 14q32.3 and PI to a more proximal position between 14q24.3 and 14q32.1. The immunoglobulin genes are in a chromosome region noted for its high frequency of breaks associated with chromosome rearrangement, occurring both spontaneously in cultured lymphocytes and in certain malignancies.

By in situ hybridization, Schroeder et al. (1985) narrowed the assignment of the PI locus to 14q31-q32. Turleau et al. (1984) studied a patient with an interstitial deletion of 14q and assigned the PI locus to 14q32.1 by exclusion mapping. In a similar patient with an interstitial deletion of 14q, Yamamoto et al. (1986) confirmed the assignment to 14q32.1. By the dosage principle, the level of alpha-1-antitrypsin in the patient was only about half of that in his parents and in controls.

Sefton et al. (1989) used pulsed field gel electrophoresis to demonstrate that the genes encoding alpha-1-antitrypsin and alpha-1-antichymotrypsin (AACT, SERPINA3; 107280) are approximately 220 kb apart and oriented in opposite directions.

Molecular studies of a ring chromosome 14 showed that the IGH and D14S1 loci were missing, whereas the PI locus was present (Keyeux et al., 1989). Thus, PI is proximal to the other 2 loci, a conclusion that was supported by much earlier data. A noncoding alpha-1-antitrypsin-like gene (PIL; 107410) is located 12 kb 3-prime of the AAT gene. Billingsley et al. (1989) found that this gene and the AAT and AACT genes are carried by a single 550-kb NarI fragment. Also see Billingsley et al. (1993).

By in situ hybridization, Ledbetter et al. (1987) localized the AAT locus to mouse chromosome 12.


Gene Function

Dycaico et al. (1988) established transgenic mouse lineages that carried the normal (M) (see 107400.0001) or mutant (Z) (107400.0011) alleles of the human AAT gene. All expressed the human protein in liver, cartilage, gut, kidneys, lymphoid macrophages, and thymus. The human M-allele protein was secreted normally into the serum. However, the human Z-allele protein accumulated in several cell types, particularly in hepatocytes, and was found in serum in concentrations 10 times lower than the M-allele protein. Mice in one lineage carrying the Z allele displayed significant runting in the neonatal period and had developed abnormalities in the liver with accumulation of human Z protein in diastase-resistant cytoplasmic globules that stained with periodic acid-Schiff reaction (PAS).

The major physiologic substrate of alpha-1-antitrypsin is elastase, particularly in the lower respiratory tract (Cox, 2001).

AAT is an acute-phase reactant in that serum levels are increased with inflammation, trauma, and pregnancy (Cox, 1989).

Alpha-1-Antitrypsin Deficiency

Deficiency of alpha-1-antitrypsin (613490) is primarily associated with the risk of emphysema and liver disease; see MOLECULAR GENETICS.

Role in Twinning

Lieberman et al. (1979) found an increased frequency of heterozygosity for antitrypsin deficiency in twins and parents of twins. They concluded that 'increased' fertility and twinning may be heterozygous advantages for antitrypsin deficiency. Clark and Martin (1982) found that the frequency of the S allele (107400.0013) in mothers of dizygotic twins (0.088) was double that in controls (0.044). The frequency of S in the parents of monozygotic twins and in fathers of DZ twins was no higher than in controls. Normal frequencies were observed for the Z allele (107400.0011). No fertility indices other than twinning itself were available. To study relationships between Pi types, fertility, and twinning, Boomsma et al. (1992) studied 90 DZ and 70 MZ Dutch twin pairs and their parents. They found that mothers of dizygotic twins had frequencies of the S and Z alleles that were 3 times higher than those in a control sample. Mothers of identical twins also had a higher frequency of S than controls. The S allele may thus both increase ovulation rate and enhance the success of multiple pregnancies.

Role in Human Immunodeficiency Virus-1 Infection

Bristow (2001) found that decreased human immunodeficiency virus (HIV) infectivity correlated significantly with decreased cell surface expression of leukocyte (neutrophil) elastase (HLE) on monocytes but not lymphocytes. Decreased levels of PI correlated with increased cell surface HLE expression and increased HIV infectivity.

Bristow et al. (2001) showed that decreased HIV viral load correlated with decreased circulating PI. Furthermore, asymptomatic patients manifested deficient levels of active PI. Bristow et al. (2001) noted that deficient levels of PI lead to degenerative lung diseases and suggested that preventing PI deficiency may prevent HIV-associated pathophysiology.

Using subclones of monocytic cell lines, Bristow et al. (2003) showed that HLE localized to the cell surface, but not granules, of HIV-1-permissive clones, and to the granules, but not the cell surface, of HIV-1-nonpermissive clones. Stimulation of nonpermissive clones with lipopolysaccharide and LBP (151990), followed by exogenous PI, induced cell surface HLE expression, resulting in susceptibility to HIV infection. PI appeared to promote HIV coreceptor colocalization with surface HLE, thus permitting HIV infectivity.

Shapiro et al. (2001) showed that, at physiologic concentrations, AAT and CE-2072, a synthetic inhibitor of neutrophil elastase and proteinase-3 (PRTN3; 177020), inhibited HIV-1 production in chronically infected monocytic cell lines, in fresh blood mononuclear cells infected after an activation step, and in permissive HeLa cells. EMSA analysis indicated that AAT suppressed activation of the HIV-1-inducing transcription factor NFKB (see 164011). In 5 individuals with the Z-type AAT mutation (glu342lys; 107400.0011), HIV-1 p24 antigen increased more than 6-fold in whole blood after infection with a monocyte-tropic HIV strain. In contrast, there was no significant increase in blood obtained from healthy volunteers.

By screening a peptide library generated from hemofiltrate, Munch et al. (2007) identified a 20-amino acid peptide from the C-proximal region of alpha-1-antitrypsin, designated virus-inhibitory peptide (VIRIP), as the most potent inhibitor of multiple HIV-1 strains, including those resistant to antiviral drugs. Changes in some VIRIP residues increased its antiviral potency 100-fold. VIRIP blocked HIV-1 entry by interacting with the virus gp41 fusion peptide. Munch et al. (2007) proposed that VIRIP may affect disease progression in HIV-1-infected individuals.

NET Inhibitory Peptides

Neonatal neutrophils fail to form neutrophil extracellular traps (NETs) due to circulating NET inhibitory peptides (NIPs), which are cleavage fragments of A1AT. Using immunofluorescence assays, Campbell et al. (2021) showed that human placenta from both term and preterm pregnancies secreted serine protease A1 (HTRA1; 602194) into fetal circulation. Plasma HTRA1 levels were reduced after delivery, and decreased HTRA1 plasma levels were associated with decreased levels of NIPs. Placental HTRA1 cleaved A1AT after amino acid 382 to generate a C-terminal cleavage fragment of A1AT, termed A1ATM383S-CF, that could inhibit NET formation in vitro. Through NET inhibition, A1ATM383S-CF decreased bacterial killing, but it maintained other key neutrophil activities in vitro. In vivo analysis with wildtype mice showed that mouse placenta also secreted Htra1, and placental Htra1 cleaved A1at to generate A1atM383S-CF and inhibit NET formation by neonatal neutrophils. Analysis with Htra1 -/- and wildtype mice revealed that inhibition of NET formation during experimental neonatal sepsis improved survival.


Nomenclature

Fagerhol (1968) suggested that the system of inherited AAT variants be called Pi for protease inhibitor. Cox (1978) reported the recommendations of a workshop on PI nomenclature.


Molecular Genetics

Many electrophoretic variants of serum alpha-1-antitrypsin have been described, beginning with those reported by Axelsson and Laurell (1965). Kueppers and Bearn (1967) studied an Italian family with multiple members heterozygous for an electrophoretic variant that could not be distinguished from that which Axelsson and Laurell (1965) found in a Swedish family.

About 30 variants of alpha-1-antitrypsin had been described by 1981 (Hug et al., 1981). The alleles were given symbols according to the relative electrophoretic mobility of the allele product.

Cox et al. (1987) studied RFLPs associated with the AAT gene. They gave information on extensive variability expressed by the polymorphic information content (PIC) as proposed by Botstein et al. (1980). PI types and M subtypes tended to be associated with specific RFLP haplotypes.

Nukiwa et al. (1988) indicated that approximately 75 AAT alleles had been identified at the protein and/or gene level.

Roychoudhury and Nei (1988) tabulated worldwide gene frequencies for allelic variants M (M1, M2, M3, M4), S, Z, F, I, and V. Cox (1989) and Crystal (1989) reviewed the variants, 'normal' and pathologic, of the PI gene.

'Normal' Alleles

Crystal (1989) listed 10 normal AAT alleles that had been sequenced (107400.0001-107400.0010).

Nukiwa et al. (1988) stated that the most common alleles are the 2 forms of M1, that with valine at position 213 (M1V; 107400.0002) and that with alanine at position 213 (M1A; 107400.0001).

'Risk' Alleles

Crystal (1989) divided AAT 'at risk' alleles into 'deficiency' alleles and 'null' alleles. He stated that except for the rare Pittsburgh allele (107400.0026), which is associated with a bleeding disorder, only those phenotypes comprising 2 'at risk' alleles place the individual at risk for development of disease. Alleles in the 'at risk' class are found almost exclusively among Caucasians of European descent, with the highest frequency in northern Europe. Blacks and Asians are rarely affected.

The most common AAT deficiency allele is the Z allele (glu342-to lys; 107400.0011), which occurs on an M1A (ala213; 107400.0001) haplotype background (Nukiwa et al., 1986). The homozygous ZZ phenotype is associated with a high risk of both emphysema and liver disease. The Z allele reaches polymorphic frequencies in Caucasians and is rare or absent in Asians and blacks (DeCroo et al., 1991; Hutchison, 1998).

Clark et al. (1982) reported the cases of 2 brothers with Weber-Christian panniculitis and the AAT Z phenotype. A younger brother had the Z phenotype without Weber-Christian disease. Along with several earlier reported cases, these observations establish a relationship.

Another common AAT deficiency allele is the S allele (glu264-to-val; 107400.0013), which occurs on an M1V (val213; 107400.0002) haplotype background. Pi*S homozygotes are at no risk of emphysema, but compound heterozygotes with a Z or a null allele have a mildly increased risk (Curiel et al., 1989). The S allele reaches polymorphic frequencies in Caucasians and is rare or absent in Asians and blacks. It is not associated with liver disease.

Other rare deficiency AAT alleles may result in increased risk for both liver and lung disease (e.g., Pi M(Malton); 107400.0012) or only emphysema (e.g., Pi M(Procida); 107400.0016). Some of the rare deficiency alleles have been found in Japanese (e.g., Pi S(Iiyama); 107400.0039).

Null AAT alleles are rare but have been found in all populations. Garver et al. (1986) investigated the molecular basis of the Pi null-null AAT phenotype. The gene appeared to be intact without discernible deletion or other structural abnormality, yet there was no detectable mRNA produced. The 5-prime promoter region also appeared to be normal. No evidence of hypermethylation of cytosine nucleotides in the promoter region was detected. The defect may be comparable to that in some forms of thalassemia in which a change, at a splicing site, for example, may lead to greatly reduced mRNA production. The null-null phenotype is accompanied by emphysema as is the ZZ and SZ phenotypes but an important difference is that cirrhosis and liver disease do not occur with the null-null phenotype; there is no abnormal antitrypsin produced that is excreted with difficulty from the cells of synthesis.

Nukiwa et al. (1987) identified a null form of alpha-1-antitrypsin resulting from a frameshift causing a stop codon to be formed approximately 44% from the N terminus of the precursor protein (Null(Granite Falls); 107400.0020). Although the molecular basis of antitrypsin deficiency was quite different from that in the Z haplotype, the phenotypic consequences were similar: severe deficiency associated with high risk of emphysema.

Seixas et al. (2002) reported 2 null alleles of the PI gene in Portuguese patients with emphysema. These alleles were associated with total lack of circulating protein as indicated by the absence of isoelectric focusing banding patterns. One of the alleles, designated Q0(Ourem), was identical to Q0(Mattawa) on an M3 normal background (107400.0022). The second allele, Q0(Porto), had a novel mutation which restricted mononuclear phagocyte transcripts to mRNA species resulting from the direct splice of exon IA to exon II. The absence of this normal splice alternative in the liver, where PI is primarily synthesized, provided a basis for the pathogenic effects of this mutation.

PI Pittsburgh

The PI Pittsburgh allele (met358-to-arg; 107400.0026), which occurs at the AAT active site, is an example of a mutation leading to altered function of the gene product. AAT becomes a potent inhibitor of thrombin and factor XI rather than of elastase. The mutation results in a bleeding disorder (Lewis et al., 1978; Owen et al., 1983).

SERPINA1 Haplotypes Associated with Chronic Obstructive Pulmonary Disease

While cigarette smoking is a major cause of COPD (see 606963), only 15% of smokers develop the disease, indicating major genetic influences. The most widely recognized candidate gene in COPD is SERPINA1, although it has been suggested that SERPINA3 (107280) may also play a role. Chappell et al. (2006) identified 15 single-nucleotide polymorphism (SNP) haplotype tags from high-density SNP maps of the 2 genes and evaluated these SNPs in the largest case-control genetic study of COPD conducted to that time. For SERPINA1, 6 newly identified haplotypes with a common backbone of 5 SNPs were found to increase the risk of disease by 6- to 50-fold, the highest risk of COPD that had been reported. In contrast, no haplotype associations for SERPINA3 were identified.


Population Genetics

DeCroo et al. (1991) studied the frequency of alpha-1-antitrypsin alleles in US whites, US blacks, and African blacks (living in Nigeria). While the PI*S allele was present at a polymorphic level in US whites, it was present only sporadically in US blacks and completely absent in African blacks. The PI*Z allele was not detected in the black populations tested. DeCroo et al. (1991) used the PI allele frequency data to calculate white admixture in US blacks. The average white admixture estimate in US blacks, based on all PI alleles, was about 13%. This value was about 24% when only the S and Z alleles were used.

Studies of the distribution of the S and Z alleles in Europe demonstrated that they occur mainly among those of European stock. Hutchison (1998) found that the frequency of the gene for PiZ is highest on the northwestern seaboard of the continent and that the mutation seems to have arisen in southern Scandinavia. The distribution of PiS is quite different: the gene frequency is highest in the Iberian peninsula and the mutation is likely to have arisen in that region.

By means of a metaanalysis of 43 studies, Blanco et al. (2001) analyzed the distribution of the PI*S and PI*Z alleles in countries outside Europe and compared them with data from Europe.


Animal Model

The pallid (pa) (604310) mouse develops emphysema late in life. Martorana et al. (1993) demonstrated that pallid mice have markedly reduced levels of serum alpha-1-antitrypsin associated with severe deficiency in serum elastase inhibitory capacity. However, they have normal alpha-1-antitrypsin mRNA levels in the liver.

Green et al. (2003) showed that Drosophila 'necrotic' (nec) mutations can mimic alpha-1-antitrypsin deficiency. They identified 2 nec mutations homologous to an antithrombin point mutation that is responsible for neonatal thrombosis. Transgenic flies carrying an amino acid substitution equivalent to that found in S(Iiyama) variant antitrypsin (107400.0039) failed to complement nec-null mutations and demonstrated a dominant temperature-dependent inactivation of the wildtype nec allele. Green et al. (2003) concluded that the Drosophila nec system can be used as a powerful system to study serpin polymerization in vivo.

Van Pel et al. (2006) reported that IL8 (146930)- and GCSF (CSF3; 138970)-induced mobilization of hematopoietic stem cells (HSCs) and hematopoietic progenitor cells (HPCs) in mice was completely inhibited by total body irradiation (TBI). They found that TBI increased expression of Serpina1 mRNA and protein, which inhibited elastase activity. Inhibition of HSC/HPC mobilization in irradiated mice could be reversed by anti-Serpina1. Furthermore, injection of Serpina1, but not heat-inactivated Serpina1, prior to administration of IL8 inhibited HSC/HPC mobilization. Van Pel et al. (2006) concluded that low-dose TBI induces Serpina1 in bone marrow and inhibits HSC/HPC mobilization, and they hypothesized that cytokine-induced HSC/HPC mobilization is determined by a critical balance between serine proteases and their inhibitors.

Kamimoto et al. (2006) bred the PiZ mouse onto the GFP-LC3 background to monitor autophagy.

Hidvegi et al. (2010) demonstrated that the autophagy-enhancing drug carbamazepine decreased the hepatic load of mutant alpha-1-antitrypsin Z (ATZ) protein and hepatic fibrosis in a mouse model of AAT deficiency-associated liver disease. The mouse used was the PiZ mouse, developed by Dycaico et al. (1988), in which the human ATZ gene is a transgene. Although the PiZ mouse has normal circulating levels of endogenous murine alpha-1-antitrypsin, it is a robust model of liver disease associated with AAT deficiency, as characterized by intrahepatocytic ATZ-containing globules, inflammation, and increased regenerative activity, dysplasia, and fibrosis. Hidvegi et al. (2010) concluded that their results in this animal model provided a basis for testing carbamazepine, which has an extensive clinical safety profile in patients with AAT deficiency (613490) and also provided proof of principle for therapeutic use of autophagy enhancers.


History

The polymorphism of prealbumin described by Fagerhol and Braend (1965) was shown by Fagerhol and Laurell (1967) to be the same as the alpha-1-antitrypsin polymorphism.

A possible heterogeneity in recombination frequency between Pi variants believed to be allelic was reported by Gedde-Dahl et al. (1972): Pi(Z) had less recombination with Gm than Pi(non-Z). Gedde-Dahl et al. (1975) gave further data on the Gm-Pi linkage. They considered heterogeneity of recombination fraction among males of different Pi types to be likely. The major difference seemed to be between the Pi(Z) and other alleles. Possible explanations included a chromosomal deletion, inversion or locus regulating recombination in linkage disequilibrium with the Pi locus. Gedde-Dahl et al. (1981) showed that the allele-specific heterogeneity of Gm-Pi linkage is attributable to 'reduced' recombination in Z-allele heterozygotes. They found an equal sex ratio for Pi 'non-Z' variants, as opposed to a 1:2 male-female ratio for 'Z' families. The location of Gm and Pi on 6p was excluded by Bender et al. (1979).

Babron et al. (1990) confirmed a previous finding that the presence of the Pi Z allele tends to decrease the recombination rate between the GM (147100) and PI loci. This decrease appeared to be similar in both sexes and not unique in males as previously noted. The results suggested a possible linkage disequilibrium between the Pi Z allele and a large inversion between the GM and PI loci.

Carrell (1986) cited evidence for the existence of 2 genes coding for alpha-1-antitrypsin, although the plasma findings were compatible with expression of the alleles at a single locus.


ALLELIC VARIANTS 40 Selected Examples):

.0001   PI M1-ALA213

PI, M1A
SERPINA1, ALA213
SNP: rs6647, gnomAD: rs6647, ClinVar: RCV000019553, RCV000019554, RCV000151834, RCV000406073, RCV001701482, RCV002362589

M1A, a normal variant, is believed to be the 'oldest' human PI allele, with the other common normal alleles M1V (107400.0002), M2 (107400.0003), and M3 (107400.0004) derived from M1A by single base substitutions. M2 is derived from M3; it has the same amino acid difference that distinguishes M3 from M1V but a second substitution in addition. The 4 common normal alleles are considered the 'base' from which all the other alleles are derived (see Fig. 4 in Crystal, 1989). The M1A allele has a frequency of 0.20-0.23 in US Caucasians.


.0002   PI M1-VAL213

PI, M1V
SERPINA1, ALA213
SNP: rs6647, gnomAD: rs6647, ClinVar: RCV000019555, RCV000019556, RCV002362590

This normal allele has a frequency of 0.44-0.49 in US Caucasians.


.0003   PI M2

SERPINA1, ARG101HIS ON M3
SNP: rs709932, gnomAD: rs709932, ClinVar: RCV000019557, RCV000019559, RCV000155576, RCV000310904, RCV001636605, RCV002345248

M2, which has a frequency of 0.10-0.11 in US Caucasians (Cox, 1989), was studied by Nukiwa et al. (1988), who found that its coding exons are identical to those of the more frequent form of M1 (val213) except for 2 bases: a change in codon 101 from CGT to CAT, leading to an amino acid change of arginine to histidine; and a change in codon 376 from GAA to GAC, resulting in an amino acid change from glutamic acid to aspartic acid. Since 2 mutations separate these 2 common alleles, Nukiwa et al. (1988) suggested that another AAT variant (presumably M3) was an intermediate in their evolution.


.0004   PI M3

SERPINA1, GLU376ASP ON M1V
SNP: rs1303, gnomAD: rs1303, ClinVar: RCV000019558, RCV000155574, RCV000380179, RCV001650837, RCV002345249

This normal variant allele has a frequency of 0.14-0.19 among US Caucasians. Graham et al. (1990) identified a single nucleotide difference between M1 (val213) and M3: a transversion in codon 376 from GAA(glu) to GAC(asp).


.0005   PI M4

SERPINA1, ARG101HIS ON M1V
ClinVar: RCV000019557, RCV000019559, RCV000155576, RCV000310904, RCV001636605, RCV002345248

M4, an uncommon normal allele, is likely derived by single substitution from M1V; however, it has the same mutation that changed M2 to M3, and thus it is possible that M4 derived from M3 (or vice versa).


.0006   PI B(ALHAMBRA)

SERPINA1, ASP-LYS
ClinVar: RCV000019560

Yoshida et al. (1979) found 2 amino acid substitutions in the rare antitrypsin variant PiB Alhambra. One substitution was asp for lys at an unknown location (Crystal, 1989).


.0007   PI F

SERPINA1, ARG223CYS ON M1V
SNP: rs28929470, gnomAD: rs28929470, ClinVar: RCV000019561, RCV000148879, RCV000151833, RCV000205893, RCV000622899, RCV000727618

This rare 'normal' allele has a CGT-to-TGT change in codon 223 (Crystal, 1989; Okayama et al., 1991).


.0008   PI P(ST. ALBANS)

SERPINA1, ASP341ASN ON M1V
SNP: rs143370956, rs28929471, gnomAD: rs143370956, ClinVar: RCV000019562, RCV000148874, RCV000178751

In addition to a GAC-to-AAC change in codon 341, this rare 'normal' allele has a 'silent' asp256 (GAT)-to-asp256 (GAC) change. The rare P-family of AAT variants is defined by the position of migration of the protein on isoelectric focusing (IEF) of serum between the common M and S variants. The P(St. Albans) allele is associated with normal serum levels of AAT, whereas the P(Lowell) allele (107400.0019) is associated with reduced levels. Holmes et al. (1990) described the DNA change underlying both of these variants.


.0009   PI X

SERPINA1, GLU204LYS ON M1V
SNP: rs199422208, gnomAD: rs199422208, ClinVar: RCV000019563, RCV000512615, RCV002260969

This rare 'normal' allele has been sequenced only at the level of the protein (Crystal, 1989).


.0010   PI CHRISTCHURCH

SERPINA1, GLU363LYS
SNP: rs121912712, gnomAD: rs121912712, ClinVar: RCV000019564, RCV000201867, RCV000597262, RCV003964806

Brennan and Carrell (1986) characterized antitrypsin Christchurch, which shows a substitution of lysine for glutamic acid at position 363. Although electrophoretic mobility of the mutant protein was abnormal, no functional abnormality of the protein was detected. The base PI allele, M1A or M1V, is unknown (Crystal, 1989).


.0011   PI Z

SERPINA1, GLU342LYS ON M1A
SNP: rs28929474, gnomAD: rs28929474, ClinVar: RCV000019567, RCV000019594, RCV000019595, RCV000148877, RCV000194811, RCV000255454, RCV000623762, RCV000768543, RCV001195107, RCV002054450, RCV002251912, RCV002276567, RCV002466247, RCV003415721

This is the most frequent allele leading to a high risk of emphysema (and liver disease) in the homozygote; the allele frequency is 0.01-0.02 in US Caucasians (Crystal, 1989). Nukiwa et al. (1986) demonstrated the val213-to-ala substitution (here symbolized M1A) in PI*Z in addition to the disease-producing glu342-to-lys mutation. Ala213 was found in all of 40 Z haplotypes, using synthetic oligonucleotide gene probes directed toward the mutated exon 3 sequences in the Z gene. Furthermore, the exon 3 mutation eliminated a BstEII restriction endonuclease site, allowing rapid identification of the change in genomic DNA. Surprisingly, only 23% of the M1 haplotypes were found to be BstEII site negative. The new form of M1, i.e., M1(ala213), is identical to M1 but has an isoelectric focusing 'silent' amino acid substitution. M1 has a frequency of 68 to 76%; M2, 14 to 20%; and M3, 10 to 12%. The Z gene represents 1 to 2% of all alpha-1-antitrypsin haplotypes.

Using 2 genomic probes extending into the 5-prime and 3-prime flanking regions, respectively, Cox et al. (1985) identified 8 polymorphic restriction sites for the PI gene. Extensive linkage disequilibrium was found with the PI Z allele throughout the probe region, but not with the normal PI M allele. The Z allele occurred mainly with one haplotype, indicating a single, relatively recent origin in Caucasians. This was an individual who lived in northern Europe some 6,000 years ago. Since then, the variant has spread through Europe with a frequency gradient extending from north to south: 5% of Scandinavians, 4% of Britons, 1 to 2% of southern Europeans, and 3% of the heterogeneous white population in the United States are MZ heterozygotes. Curiously, there is a reciprocal distribution of the S variant form: 10% in southern Europe to 5% in the north. As a general rule then, 1 in 10 persons of European origin will be heterozygous for either the S or Z variant, i.e., MZ or MS (Carrell, 1986). Kawakami et al. (1981) cited 2 studies in which no Pi Z was found among 965 healthy Japanese and 183 Japanese with pulmonary diseases. This is to be compared with a frequency of 1.6% for Pi Z among Norwegians.

Crystallographic analysis of alpha-1-antitrypsin predicts that in the normal protein a negatively charged glu342 is adjacent to a positively charged lys290. Thus, the glu342-to-lys Z mutation causes the loss of a normal salt bridge, resulting in intracellular aggregation of the Z molecule. Brantly et al. (1988) predicted that a second mutation that changed the positively charged lys290 to a negatively charged glu290 would correct the secretion defect. They demonstrated that such was the case: when the second mutation was added to the Z-type cDNA, the resulting gene directed the synthesis and secretion of AAT similar to that directed by the normal AAT cDNA in an in vitro eukaryotic expression system. In general it may be possible to correct human hereditary disease by inserting an additional mutation in the gene.

By analyzing nonrecombinant SNPs of 21 Latvian and 65 Swedish heterozygous and homozygous PI Z allele carriers and 113 healthy Latvian controls, Lace et al. (2008) estimated the age of the PI Z mutation to be 2,902 years in Latvia and 2,362 years in Sweden. The SNPs showed a high degree of similarity between the 2 populations, indicating a common ancestor.

Approximately 3 to 5% of patients with cystic fibrosis (CF; 219700) develop severe liver disease defined as cirrhosis with portal hypertension. Bartlett et al. (2009) performed a 2-stage case control study enrolling patients with CF and severe liver disease with portal hypertension from 63 CF centers in the United States as well as 32 in Canada and 18 outside of North America. In the first stage, 124 patients with CF and severe liver disease, enrolled between January 1999 and December 2004, and 843 control patients without CF-related liver disease (all assessed at greater than 15 years of age) were studied by genotyping 9 polymorphisms in 5 genes previously studied as modifiers of liver disease in CF. In the second stage, the 2 genes that were positive from the first stage were tested in an additional 136 patients with CF-related liver disease enrolled between January 2005 and February 2007, and in 1,088 with no CF-related liver disease. The combined analysis of the initial and replication studies by logistic regression showed CF-related liver disease to be associated with the SERPINA1 Z allele (odds ratio = 5.04; 95% confidence interval 2.88-8.83; p = 1.5 x 10(-8)). Bartlett et al. (2009) concluded that the SERPINA1 Z allele is a risk factor for liver disease in CF. Patients carrying the Z allele are at greater risk (odds ratio = approximately 5) of developing severe liver disease with portal hypertension.

Using several markers of ER stress response, Kelly et al. (2009) found that expression of AAT with the Z mutation, which they called ZAAT, resulted in ER stress in transfected HepG2 cells. Coexpression of the ER stress response selenoprotein SEPS1 (607918) relieved ER stress caused by ZAAT expression or by exposure to tunicamycin, a known ER stressor. Supplementation of cells with selenium augmented the activity of SEPS1. Selenium supplementation also increased endogenous SEPS1 expression and reduced ER stress.


.0012   PI M(MALTON)

SERPINA1, PHE52DEL ON M2
SNP: rs775982338, gnomAD: rs775982338, ClinVar: RCV000019568, RCV000019572, RCV000262058, RCV000594562, RCV003417992

Liver disease, as well as emphysema, has been described in patients with the rare PI*M(Malton) allele. Fraizer et al. (1989) studied the molecular defect in M(Malton), a deficiency allele which, like the Z allele, is associated with hepatocyte inclusions and impairs secretion. They found that the M(Malton) allele contains a deletion of the codon for 1 of the 2 adjacent phenylalanine residues (amino acid 51 or 52 of the mature protein). Judging from the haplotype data, the M(Malton) mutation must have derived from the normal M2 allele. Deletion of the 1 amino acid would be expected to shorten 1 strand of the beta-sheet, B6, apparently preventing normal processing and secretion. Curiel et al. (1989) also showed that the M(Malton) allele differs from the normal M2 allele by deletion of the entire codon (TTC) for residue phe52. They demonstrated abnormal intracellular accumulation of newly synthesized AAT protein in a homozygote who also showed, on liver biopsy, inflammation, mild fibrosis, and intrahepatocyte accumulation of the protein. Furthermore, Curiel et al. (1989) showed by retroviral gene transfer of AAT cDNA with the M(Malton) phe52 deletion into murine cells that abnormal accumulation of the newly synthesized protein occurred. This provides further evidence that abnormal intrahepatocyte AAT accumulation is responsible for the liver injury. By means of gene amplification and direct DNA sequencing, Graham et al. (1989) identified the same mutation, pointing out that it could be either phenylalanine-51 or phenylalanine-52 that is deleted.


.0013   PI S

SERPINA1, GLU264VAL ON M1V
SNP: rs17580, gnomAD: rs17580, ClinVar: RCV000019569, RCV000148878, RCV000177031, RCV000508742, RCV000508836, RCV000762932, RCV000768544, RCV000991136, RCV001195102, RCV002371777, RCV002466248, RCV003415722

Owen and Carrell (1976) and Yoshida et al. (1977) found substitution of valine for glutamic acid at position 264 in the S variant of alpha-1-antitrypsin. See Long et al. (1984).

Curiel et al. (1989) concluded that the S-type AAT protein is degraded intracellularly before secretion. PI*S homozygotes are at no risk of emphysema, but compound heterozygotes with Z or a null allele have a mildly increased risk. Because of the high frequency of the PI*S allele (0.02-0.04 in US Caucasians), such compound heterozygotes are relatively frequent.


.0014   PI M(HEERLEN)

SERPINA1, PRO369LEU ON M1A
SNP: rs199422209, gnomAD: rs199422209, ClinVar: RCV000019565, RCV000409001, RCV000727230, RCV002336088, RCV003944831

Hofker et al. (1989) demonstrated the molecular defect in the PI gene of a patient with a serum level of only 5 mg/100 ml and a PI M-like phenotype, designated PI M(Heerlen). They demonstrated a substitution of leucine for proline at codon 369, which resulted from a C-to-T mutation in exon 5. Otherwise the nucleotide sequence of the exons, intron/exon junctions, and a part of the promoter region was similar to that of a PI M1(ala213) gene. Kalsheker et al. (1992) described a family with Pi M(Heerlen) and commented on the difficulties of diagnosis of rare PI (null) or Q0 variants.


.0015   PI M(MINERAL SPRINGS)

SERPINA1, GLY67GLU ON M1A
SNP: rs28931568, ClinVar: RCV000019566, RCV000201855

This mutation, which causes AAT deficiency and emphysema, is unique among antitrypsin mutations in that it was observed in a black family, whereas most mutations causing AAT deficiency are confined to Caucasian populations of European descent. The index case was homozygous. A GGG-to-GAG change in codon 67 led to substitution of glutamic acid for glycine (Curiel et al., 1990). Curiel et al. (1990) showed that this mutation caused reduced AAT secretion on the basis of aberrant posttranslational biosynthesis by a mechanism distinct from that associated with the Z allele, whereby intracellular aggregation of the mutant protein is responsible for the secretory defect. Furthermore, the M(Mineral Springs) mutation markedly affected the ability of the protein that did reach the circulation to inhibit neutrophil elastase. Homozygotes have a high risk of emphysema (Crystal, 1989).


.0016   PI M(PROCIDA)

SERPINA1, LEU41PRO ON M1V
SNP: rs28931569, gnomAD: rs28931569, ClinVar: RCV000019571, RCV000201848, RCV000729804, RCV001807004

Takahashi et al. (1988) showed that M(Procida) has a substitution of proline for leucine at position 41, resulting from a change of codon CTG to CCG. The rare mutant protein shows somewhat reduced catalytic activity; its concentration is low in plasma, apparently because of instability and resulting intracellular degradation before secretion. Homozygotes have a high risk of emphysema (Crystal, 1989).


.0017   PI M(NICHINAN)

SERPINA1, PHE52DEL AND GLY148ARG
SNP: rs112030253, gnomAD: rs112030253, ClinVar: RCV000019568, RCV000019572, RCV000262058, RCV000512621, RCV000594562, RCV000597384, RCV003417992

Nakamura et al. (1980) found this variant in a 42-year-old Japanese woman with neither pulmonary emphysema nor liver dysfunction. She was the product of a consanguineous marriage. Radial immunodiffusion assay showed a low level of AAT in serum (17.9 mg/dl as compared to the normal range of 190-280 mg/dl). Aggregation of AAT molecules was demonstrated histologically in hepatocytes, indicating profound reduction in the secretion of the protein. Serum AAT levels in the members of the family demonstrated that the proband was homozygous for the M(Nichinan) allele. Matsunaga et al. (1990) demonstrated that the M(Nichinan) gene is identical with the M1(val213) gene except for 2 changes: a TTC trinucleotide deletion in the codon for phenylalanine-52 and a G-A substitution by which the normal gly148(GGG) became arg148(AGG). Matsunaga et al. (1990) suggested that the gly148-to-arg change is unlikely to be the cause of the AAT deficiency because arg (not gly) is located at the corresponding position of the protein C inhibitor which belongs to the same family of serine protease. On the other hand, Matsunaga et al. (1990) suggested that deletion of phenylalanine-52 may cause the newly synthesized AAT protein to aggregate, resulting in serum AAT deficiency. They suggested that the gly148-to-arg substitution reflects the vulnerability of a CpG dinucleotide to mutation. They pointed to a number of other variant forms of AAT that were probably generated through a C-T transition. Indeed, the Z and M1(val213) genes were generated from the M1(ala213) gene by the C-T transition at the CpG dinucleotide on the antisense and the sense strands, respectively. The M2 gene was generated from the M3 gene by the same mechanism.


.0018   PI I

SERPINA1, ARG39CYS ON M1V
SNP: rs28931570, gnomAD: rs28931570, ClinVar: RCV000019575, RCV000148875, RCV000205893, RCV000431149, RCV003390693

By gene amplification and direct DNA sequencing, Graham et al. (1989) identified this mutation, CGC to TGC, in a compound heterozygote. Homozygotes are at no risk of emphysema, but compound heterozygotes with Z or a null allele have a mildly increased risk (Crystal, 1989). In 1 individual and 3 independent families, Seri et al. (1992) confirmed that the I variant resulted from a CGC (arg)-to-TGC (cys) transition at codon 39 within exon 2.


.0019   PI P(LOWELL)

PI NULL(CARDIFF)
PI Q0(CARDIFF)
SERPINA1, ASP256VAL ON M1V
SNP: rs121912714, gnomAD: rs121912714, ClinVar: RCV000019576, RCV000019577, RCV000019578, RCV000019605, RCV000148876, RCV000398063, RCV003415723

Faber et al. (1989) demonstrated that this rare allele, a cause of deficiency of alpha-1-antitrypsin, results from an A-to-T transversion in exon 3 of the gene. As a result, GAT (aspartic acid at residue 256) is converted to GTT (valine at that position). The same change was found in a total of 4 families.

By gene amplification and direct DNA sequencing, Graham et al. (1989) identified the same mutation in a variant they called Null(Cardiff). According to the tabulation by Crystal (1989), homozygotes have no risk for emphysema, but compound heterozygotes with a Z or null allele have a mildly increased risk.

By retroviral insertion of the P(Lowell) cDNA into the genome of NIH-3T3 fibroblasts, Holmes et al. (1990) demonstrated a pattern of biosynthesis of AAT consistent with the intracellular degradation of newly synthesized protein. Because serum AAT deficiency associated with other mutations resulting from intracellular degradation of the protein can be overcome by administration of estrogenlike drugs, Holmes et al. (1990) administered tamoxifen to a subject with the P(Lowell)/Z phenotype and demonstrated a 48% rise in AAT serum levels over a 5-month period, from below the threshold for protection from emphysema to a value above that threshold. Seri et al. (1992) confirmed the nature of the mutation in P(Lowell).

Hildesheim et al. (1993) demonstrated that P(Duarte) (107400.0037) has the same mutation as that in P(Lowell) but that it is on a background of the normal M4 allele (R101H; 107400.0005). Hildesheim et al. (1993) pointed out that this is an example of genetic diversity resulting from a limited repertoire of mutations on different common allelic backgrounds--a combinatorial basis for genetic diversity. A similar example is the occurrence of Creutzfeldt-Jakob disease and fatal familial insomnia as a result of the same mutation, depending on the nature of a nucleotide polymorphism at another site in the prion protein gene (PRNP; 176640.0010).


.0020   PI NULL(GRANITE FALLS)

PI Q0(GRANITE FALLS)
SERPINA1, TYR160TER ON M1A
SNP: rs267606950, gnomAD: rs267606950, ClinVar: RCV000019579, RCV000019580

The gene shows deletion of the third nucleotide in the tyr160 codon TAC, causing a frameshift with new stop codon TAG at position 160 (Nukiwa et al., 1987). Emphysema is associated with homozygosity.


.0021   PI NULL(BELLINGHAM)

PI Q0(BELLINGHAM)
SERPINA1, LYS217TER ON M1V
SNP: rs199422211, gnomAD: rs199422211, ClinVar: RCV000019581, RCV000019582, RCV000169162, RCV001528835

By cloning and sequencing the Null(Bellingham) gene (which in homozygous state is associated with early-onset emphysema), Satoh et al. (1988) demonstrated that the promoter region, coding exons, and all exon-intron junctions are normal except for a single base substitution in exon 3, which causes the normal lys217 (AAG) to become a stop codon (TAG).


.0022   PI NULL(MATTAWA)

SERPINA1, LEU353PHE ON M1V
SNP: rs28929473, gnomAD: rs28929473, ClinVar: RCV000019583

Cox and Levison (1988) reported a family in which several members manifested no detectable plasma alpha-1-antitrypsin (613490), indicating a 'null' AAT allele, which the authors designated Null Mattawa (QO-Mattawa). Curiel et al. (1989) studied 2 affected sisters in this family and found that they were compound heterozygous for 2 mutations in the AAT gene: Null(Bellingham) (107400.0021) and Null(Mattawa). Sequencing of exons 1c-5 and all exon-intron junctions of the Null(Mattawa) gene demonstrated that it was identical to the common normal M1(val213) gene except for the insertion of a single nucleotide within the coding region of exon 5, causing a 3-prime frameshift with generation of a premature stop signal at position 376. Monocytes were shown to have an mRNA transcript of normal size, and in vitro translation showed that the mRNA was translated at a normal rate but produced a truncated antitrypsin protein. Additionally, retroviral transfer of the cDNA to murine fibroblasts demonstrated no detectable intracellular or secreted protein despite the presence of Null(Mattawa) mRNA. Thus, the molecular pathophysiology of Null(Mattawa) is probably manifested at a posttranslational level. This allele is associated with high risk of emphysema.


.0023   PI NULL(PROCIDA)

PI NULL(ISOLA DI PROCIDA)
PI Q0(PROCIDA)
SERPINA1, 17-KB DEL
ClinVar: RCV000019584, RCV000203601

Of the 5 previously known representatives of the 'null' group of AAT-deficient alleles (i.e., genes incapable of producing AAT protein detectable in serum) evaluated at the gene level, all had stop codons in coding exons. Cloning and mapping of the Null(Isola di Procida) gene demonstrated deletion of a 17-kb fragment that included exons 2-5 of the AAT structural gene (Takahashi and Crystal, 1990). Sequence analysis showed a 7-bp repeat sequence both 5-prime to the deletion and at the 3-prime end of the deletion, suggesting that the mechanism of the deletion may have been a slipped mispairing. This mutation, which at first was called Null(Procida), was found in heterozygous state with the M(Procida) allele (107400.0016) reported by Takahashi et al. (1988). To avoid confusion with M(Procida), Null(Procida) was renamed Null(Isola di Procida). This mutation is associated with high risk of emphysema.


.0024   PI NULL(HONG KONG 1)

PI Q0(HONG KONG 1)
SERPINA1, 2-BP DEL, FS334TER
SNP: rs1057519610, ClinVar: RCV000019587, RCV000019588, RCV000201857

Deletion of TC from CTC codon 318 for leucine causes frameshift with stop codon TAA at position 334. Homozygosity for this allele, like other null alleles, predisposes to early-onset emphysema. See Sifers et al. (1988). This variant was initially called Null(Hong Kong) but later Null(Hong Kong-1) because a second null allele called Null(Hong Kong-2) (107400.0034) was identified in the same individual by haplotype analysis (Fraizer et al., 1990).


.0025   PI NULL(BOLTON)

PI Q0(BOLTON)
SERPINA1, 1-BP DEL
SNP: rs764325655, gnomAD: rs764325655, ClinVar: RCV000019589, RCV000019590, RCV000512620

Fraizer et al. (1989) observed a unique PI null allele. By cloning and sequencing the allele, they demonstrated deletion of a single cytosine residue (the third C in the CCC codon 362 for proline) near the active site of alpha-1-antitrypsin in exon 5 resulting in a frameshift which caused an in-frame stop codon downstream of the deletion. The stop codon led to premature termination of protein translation at amino acid 373, resulting in a truncated protein. PI Q0(Bolton) was observed in combination with PI*M(Malton) in 2 compound heterozygotes. The allele carries a high risk of emphysema.


.0026   PI PITTSBURGH

'ANTITHROMBIN' PITTSBURGH
SERPINA1, MET358ARG
SNP: rs121912713, gnomAD: rs121912713, ClinVar: RCV000019591, RCV000201860

This structure mutation in the PI gene alters its function such that it becomes an antithrombin and leads to a bleeding disorder. Alpha-1-antitrypsin and antithrombin III (107300) have a similar structure reflecting origin from a common ancestral protein some 500 million years ago. Both are inhibitors of proteolytic enzymes but have different specificities. Alpha-1-antitrypsin protects the body against released elastase, whereas AT III controls coagulation by inhibiting thrombin and other activated coagulation factors. Owen et al. (1983) described a mutation of alpha-1-antitrypsin that converts it to an antithrombin. Whereas synthesis of alpha-1-antitrypsin increases in response to trauma, AT III remains at a constant plasma concentration and requires activation by heparin. The antithrombin activity of the mutant alpha-1-antitrypsin was independent of heparin but its synthesis was stimulated by trauma. The patient was a 14-year-old boy who died in 1981 with a huge hematoma of his leg and abdomen. This was the last of a lifelong series of bleeding episodes occurring after trauma and requiring hospitalization on more than 50 occasions. Lewis et al. (1978) described the clinical picture and identified a variant 'antithrombin' which they called antithrombin Pittsburgh. It had, however, the electrophoretic and antigenic characteristics of a variant alpha-1-antitrypsin. Owen et al. (1983) showed that the variant protein has arginine at position 358, replacing the normal methionine. This finding indicated that the reactive center of alpha-1-antitrypsin is methionine 358, which acts as a 'bait' for elastase, just as the normal reactive center of AT III is arginine-393, which acts as a bait for thrombin. Neutrophils augment tissue proteolysis by the oxidative inactivation of the methionine at the reactive center of alpha-1-antitrypsin. Scott et al. (1986) and Schapira et al. (1986) found that recombinant AAT-Pittsburgh (met358-to-arg) is a potent inhibitor of plasma kallikrein and activated factor XII fragment, although it has lost its anti-elastase activity. They suggested it might have therapeutic potential in hereditary angioedema or septic shock. Vidaud et al. (1992) demonstrated that a G-to-T transversion at nucleotide 10038 is responsible for the substitution of arg for met, which converts alpha-1-antitrypsin into an arg-ser protease inhibitor (serpin) that inhibits thrombin and factor Xa more effectively than antithrombin III. They observed a 15-year-old boy who surprisingly had no bleeding history. They suggested that a large decrease in protein C concentration may account for the mild or absent bleeding tendency. The deficiency of protein C in turn was attributed to deleterious effect of the abnormal inhibitor on both intracellular processing and catabolism of protein C. In later studies, Emmerich et al. (1995) suggested that strong affinity of the mutant AAT for protein C leads in the patient of Vidaud et al. (1992) to an increased turnover and thus to a low circulating level of protein C. They proposed that in the presence of the Pittsburgh mutant protein C can be activated and is abnormally rapidly cleared. The resultant relative lack of protein C anticoagulant function may ameliorate the bleeding diathesis expected to be associated with the Pittsburgh mutation.

Wilkie (1994) discussed the molecular basis of genetic dominance and provided a useful table. He indicated altered substrate specificity as one mechanism and antithrombin Pittsburgh as a specific example.


.0027   PI V(MUNICH)

SERPINA1, ASP2ALA ON M1V
SNP: rs199422212, gnomAD: rs199422212, ClinVar: RCV000019593, RCV000512626

In an alpha-1-antitrypsin variant called V(Munich) because the major fraction focused in the 'V' region of the isoelectric focusing gel, Holmes et al. (1990) found that the molecule differs from that of the common M1V allele by a single nucleotide substitution of cytosine for adenosine, with the resultant amino acid change asp2 to ala; the codon change is GAT to GCT.


.0028   PI Z(AUGSBURG)

PI Z(TUN)
SERPINA1, GLU342LYS ON M2
ClinVar: RCV000019567, RCV000019594, RCV000019595, RCV000148877, RCV000194811, RCV000255454, RCV000623762, RCV000768543, RCV001195107, RCV002054450, RCV002251912, RCV002276567, RCV002466247, RCV003415721

Using isoelectric focusing with a narrow pH gradient, Weidinger et al. (1985) recognized a rare deficient PI variant, which they called PI Z(Augsburg). To their surprise, Faber et al. (1990) found that the sequence of the Z(Augsburg) gene showed the common PI*Z mutation (M1 glu342 GAG to Z lys342 AAG) which occurred, however, in an M2 ancestral gene. Previous findings indicated that the Z mutation had always been derived from an M1 ala213 background gene. Whitehouse et al. (1989) studied 2 sibs with mild liver abnormality who were found to be compound heterozygotes for the classical PI*Z allele and an allele that they called PI*Z(Tun). The Z(Tun) protein appeared to be deficient in the plasma to about the same degree as the Z protein. They found that the mutation was precisely the same as that in the Z allele, namely, a G-to-A transition at codon 342 resulting in the substitution of lysine for glutamic acid; however, the Z(Tun) mutation had occurred on an M2-like haplotype background rather than the M1A background. Because of its association with a unique DNA haplotype and the gene frequency estimates in populations of European origin, the Z mutation is thought to have occurred only once, about 6,000 years ago, in a North European person. The Z gene is very rare among other ethnic groups.


.0029   PI W(BETHESDA)

SERPINA1, ALA336THR ON M1A
SNP: rs1802959, gnomAD: rs1802959, ClinVar: RCV000019596, RCV000512619, RCV000734987

This variant allele, which is associated with increased risk of emphysema and liver disease, has a mutation in exon 5 where codon 336 is changed from GCT to ACT, resulting in substitution of threonine for alanine (Crystal, 1990). Holmes et al. (1990) reported that the W(Bethesda) form differs from the normal M1(ala-213) allele by a change in codon 336 from GCT to ACT. Although W(Bethesda) mRNA was translated normally in vitro, transfection of the W(Bethesda) cDNA into COS-I cells was associated with AAT secretion only 50% that of cells transfected with normal cDNA. There was no intracellular accumulation as observed with the Z allele, but reduced intracellular AAT suggested degradation of newly synthesized W(Bethesda) molecules.


.0030   PI NULL(DEVON)

PI Q0(DEVON)
PI NULL(NEWPORT)
PI Q0(NEWPORT)
SERPINA1, GLY115SER
SNP: rs11558261, gnomAD: rs11558261, ClinVar: RCV000019597, RCV000019598, RCV000019599, RCV000019600, RCV000512622

This variant, which is associated with increased risk of emphysema and liver disease, is due to a change in exon 2, resulting in substitution of serine for glycine-115 (Crystal, 1990). In a compound heterozygote carrying the common disease-producing mutation Pi Z (107400.0011), Graham et al. (1990) found a substitution of glycine-115 by serine. The mutation occurred on the background of M3. A change in codon 115 from GGC to AGC was responsible.


.0031   PI NULL(LUDWIGSHAFEN)

PI Q0(LUDWIGSHAFEN)
SERPINA1, ILE92ASN
SNP: rs28931572, ClinVar: RCV000019601, RCV000019602, RCV000201851

In this variant, which is associated with increased risk of emphysema and liver disease, a change in codon 92 from ATC to AAC in exon 2 results in substitution of asparagine for isoleucine (Crystal, 1990). This substitution of a polar for a nonpolar amino acid occurs in 1 of the alpha-helices and is predicted to disrupt the tertiary structure (Fraizer et al., 1990). Fraizer et al. (1990) identified a T-to-A substitution in a German patient.


.0032   PI Z(WREXHAM)

SERPINA1, SER-19LEU
SNP: rs140814100, gnomAD: rs140814100, ClinVar: RCV000019570, RCV000148880, RCV000596635

In a compound heterozygote with the common disease-producing PI Z mutation (107400.0011), Graham et al. (1990) found a change from TCG to TTG in codon -19, which resulted in a change from serine to leucine in the signal peptide.


.0033   MOVED TO 107400.0028


.0034   PI NULL(HONG KONG 2)

PI Q0(HONG KONG 2)
SERPINA1,
ClinVar: RCV000019573

See 107400.0024 and Fraizer et al. (1990).


.0035   PI NULL(RIEDENBURG)

SERPINA1, DEL
ClinVar: RCV000019603

Poller et al. (1991) found complete deletion of the AAT gene as the basis for PI Q0(Riedenburg). The deletion extended into the 3-prime flanking region of the gene but did not include the noncoding AAT-related gene (PIL), which is located 12 kb downstream of AAT (Hofker et al., 1988).


.0036   PI KALSHEKER-POLLER

SERPINA1, G-A, 3-PRIME UTR ENHANCER
SNP: rs11568814, gnomAD: rs11568814, ClinVar: RCV000019604, RCV000333809

Kalsheker et al. (1987) and Poller et al. (1990) reported a mutation in the 3-prime flanking sequence of the AAT gene that occurs in about 17% of patients with chronic respiratory disease. The mutation is a G-to-A nucleotide substitution in an octamer (OCT)-like sequence. Because TCGA is converted to TCAA, the mutation is detected as a restriction fragment length polymorphism with the restriction enzyme TaqI. The mutation does not appear to affect basal expression of the protein as the plasma concentration of alpha-1-antitrypsin is normal in persons who carry the mutation; however, binding and functional studies by Morgan et al. (1993) suggested that it may reduce the rise in plasma AAT concentration that occurs during inflammation. Stimulation by cytokines, such as interleukin-6 (IL6; 147620), may be lacking. Morgan et al. (1993) pointed out a precedent for such a mechanism in an unrelated gene: an enhancer element in the 3-prime flanking sequence of the erythropoietin gene increases gene expression nearly 15-fold during hypoxia.


.0037   PI P(DUARTE)

SERPINA1, ASP256VAL
ClinVar: RCV000019576, RCV000019577, RCV000019578, RCV000019605, RCV000148876, RCV000398063, RCV003415723

Hildesheim et al. (1993) demonstrated that the deficiency-producing change in the PI gene in P(Duarte) is the same as that in P(Lowell) (107400.0019). The alleles differ with respect to polymorphic nucleotides at other positions in the gene. They referred to this as genetic diversity from a limited repertoire of mutations on different common allelic backgrounds.


.0038   PI NULL(WEST)

PI Q0(WEST)
SERPINA1, IVS2DS, G-T, +1
SNP: rs751235320, gnomAD: rs751235320, ClinVar: RCV000019606, RCV000169461, RCV000593270

During routine screening of individuals applying for enrollment in the US AAT Deficiency Registry, Laubach et al. (1993) identified a patient with emphysema and a PI type heterozygous for a novel AAT null allele. The novel allele, designated PI*Q0(West), was characterized by a single G-to-T transversion at position 1 of intron 2, a highly conserved nucleotide position. This resulted in an in-frame deletion of amino acids gly164-to-lys191. This was the first splicing mutation observed in the AAT gene.


.0039   PI S(IIYAMA)

SERPINA1, SER53PHE
SNP: rs55819880, gnomAD: rs55819880, ClinVar: RCV000019608, RCV000169508

In a 32-year-old Japanese male with pulmonary emphysema, Yuasa et al. (1993) demonstrated homozygosity for a C-to-T transition at codon 53 resulting in substitution of phenylalanine for serine. They commented on the fact that, in Japanese, deficiency in null alleles at the AAT locus are extremely rare and PI*Z, which occurs at polymorphic frequencies in Caucasians, has not been reported. The only other Japanese case of AAT deficiency was that due to PI M(Nichinan) (107400.0017) reported by Matsunaga et al. (1990).


.0040   PI Z(BRISTOL)

SERPINA1, THR85MET ON M1V
SNP: rs199422213, gnomAD: rs199422213, ClinVar: RCV000019609, RCV000512618, RCV000731628

In a woman with an obstetric history of 3 perinatal deaths from fulminant liver disease and no living offspring, Lovegrove et al. (1997) found that she and her father were both heterozygotes for PI M1Z(Bristol). The Z(Bristol) protein was found to be active as a proteinase inhibitor but appeared to be deficient in the plasma to about the same degree as the S protein in MS heterozygotes. It focused on the basic side of Z and lacked the normal pattern of secondary isoforms associated with the commonly occurring AAT variants and migrated faster than normal on an SDS electrophoresis gel. The Z(Bristol) mutation was found to be a C-to-T transition at codon 85, changing ACG (thr) to ATG (met). This disrupted the N-glycosylation site starting at asn83, preventing glycosylation at residue 83 in the PI Z(Bristol) protein, and explained the protein isoelectric focusing and SDS gel electrophoresis results. An analysis of haplotypes in the propositus and her father indicated that the Z(Bristol) mutation occurred on the common M1(val213) genetic background. Of the 3 offspring with perinatal death from fulminant liver disease, 2 were by the woman's husband and 1 by an artificial insemination donor. Of the 2 offspring who were tested for the mutation, 1 had the variant and the other did not. Thus, the relationship between Z(Bristol) and fulminant liver disease in the offspring was unclear.


See Also:

Arnaud et al. (1978); Carrell et al. (1982); Chapuis-Cellier et al. (1981); Cox and Smyth (1983); Cox (1975); Cox (1980); Cox (1981); Faber et al. (1994); Faber et al. (1989); Fagerhol and Cox (1981); Fagerhol and Gedde-Dahl (1969); Fagerhol and Hauge (1968); Fagerhol and Laurell (1970); Fagerhol and Tenfjord (1968); Frants and Eriksson (1980); Gedde-Dahl et al. (1975); Graham et al. (1990); Holmes et al. (1990); Holmes et al. (1990); Hug et al. (1980); Iammarino et al. (1979); Kramps et al. (1981); Kueppers and Christopherson (1978); Lopez et al. (1964); Nukiwa et al. (1986); Owen et al. (1976); Weitkamp et al. (1978); Welch et al. (1980); Yoshida et al. (1976); Yoshida et al. (1979)

REFERENCES

  1. Arnaud, P., Galbraith, R. M., Galbraith, G. M. P., Allen, R. C., Fudenberg, H. H. A new allele of human alpha-1-antitrypsin: Pi (N Hampton). Am. J. Hum. Genet. 30: 653-659, 1978. [PubMed: 311584]

  2. Axelsson, U., Laurell, C. B. Hereditary variants of serum alpha-1-antitrypsin. Am. J. Hum. Genet. 17: 466-472, 1965. [PubMed: 4158556]

  3. Babron, M. C., Constans, J., Dugoujon, J. M., Cambon-Thomsen, A., Bonaiti-Pellie, C. The Gm-Pi linkage in 843 French families: effect of the alleles Pi Z and Pi S. Ann. Hum. Genet. 54: 107-113, 1990. [PubMed: 2382967] [Full Text: https://doi.org/10.1111/j.1469-1809.1990.tb00366.x]

  4. Bartlett, J. R., Friedman, K. J., Ling, S. C., Pace, R. G., Bell, S. C., Bourke, B., Castaldo, G., Castellani, C., Cipolli, M., Colombo, C., Colombo, J. L., Debray, D., and 22 others. Genetic modifiers of liver disease in cystic fibrosis. JAMA 302: 1076-1083, 2009. [PubMed: 19738092] [Full Text: https://doi.org/10.1001/jama.2009.1295]

  5. Bender, K., Muller, C. R., Schmidt, A., Strohmaier, U., Wienker, T. F. Linkage studies on the human Pi, Gm, GLO, and HLA genes. Hum. Genet. 49: 159-166, 1979. [PubMed: 112033] [Full Text: https://doi.org/10.1007/BF00277637]

  6. Billingsley, G. D., Walter, M. A., Cox, D. W. Physical linkage of alpha-1-antitrypsin and alpha-1-antichymotrypsin by pulsed field gel electrophoresis. (Abstract) Am. J. Hum. Genet. 45 (suppl.): A130, 1989.

  7. Billingsley, G. D., Walter, M. A., Hammond, G. L., Cox, D. W. Physical mapping of four serpin genes: alpha-1-antitrypsin, alpha-1-antichymotrypsin, corticosteroid-binding globulin, and protein C inhibitor, within a 280-kb region on chromosome 14q32.1. Am. J. Hum. Genet. 52: 343-353, 1993. [PubMed: 8381582]

  8. Blanco, I., Bustillo, E. F., Rodriguez, M. C. Distribution of alpha-1-antitrypsin PI S and PI Z frequencies in countries outside Europe: a meta-analysis. Clin. Genet. 60: 431-441, 2001. [PubMed: 11846735] [Full Text: https://doi.org/10.1034/j.1399-0004.2001.600605.x]

  9. Boomsma, D. I., Frants, R. R., Bank, R. A., Martin, N. G. Protease inhibitor (Pi) locus, fertility and twinning. Hum. Genet. 89: 329-332, 1992. [PubMed: 1601424] [Full Text: https://doi.org/10.1007/BF00220552]

  10. Botstein, D., White, R. L., Skolnick, M., Davis, R. W. Construction of a genetic linkage map in man using restriction fragment length polymorphisms. Am. J. Hum. Genet. 32: 314-331, 1980. [PubMed: 6247908]

  11. Brantly, M., Courtney, M., Crystal, R. G. Repair of the secretion defect in the Z form of alpha-1-antitrypsin by addition of a second mutation. Science 242: 1700-1702, 1988. [PubMed: 2904702] [Full Text: https://doi.org/10.1126/science.2904702]

  12. Brennan, S. O., Carrell, R. W. Alpha-1-antitrypsin Christchurch, 363 glu-to-lys: mutation at the P-prime-5 position does not affect inhibitory activity. Biochim. Biophys. Acta 873: 13-19, 1986. [PubMed: 3527273] [Full Text: https://doi.org/10.1016/0167-4838(86)90183-4]

  13. Bristow, C. L., Mercatante, D. R., Kole, R. HIV-1 preferentially binds receptors copatched with cell-surface elastase. Blood 102: 4479-4486, 2003. [PubMed: 12933574] [Full Text: https://doi.org/10.1182/blood-2003-05-1635]

  14. Bristow, C. L., Patel, H., Arnold, R. R. Self antigen prognostic for human immunodeficiency virus disease progression. Clin. Diagn. Lab. Immun. 8: 937-942, 2001. [PubMed: 11527807] [Full Text: https://doi.org/10.1128/CDLI.8.5.937-942.2001]

  15. Bristow, C. L. Slow human immunodeficiency virus (HIV) infectivity correlated with low HIV coreceptor levels. Clin. Diagn. Lab. Immun. 8: 932-936, 2001. [PubMed: 11527806] [Full Text: https://doi.org/10.1128/CDLI.8.5.932-936.2001]

  16. Campbell, R. A., Campbell, H. D., Bircher, J. S., de Araujo, C. V., Denorme, F., Crandell, J. L., Rustad, J. L., Monts, J., Cody, M. J., Kosaka, Y., Yost, C. C. Placental HTRA1 cleaves alpha-1-antitrypsin to generate a NET-inhibitory peptide. Blood 138: 977-988, 2021. [PubMed: 34192300] [Full Text: https://doi.org/10.1182/blood.2020009021]

  17. Carrell, R. W., Jeppsson, J.-O., Laurell, C.-B., Brennan, S. O., Owen, M. C., Vaughan, L., Boswell, D. R. Structure and variation of human alpha-1-antitrypsin. Nature 298: 329-334, 1982. [PubMed: 7045697] [Full Text: https://doi.org/10.1038/298329a0]

  18. Carrell, R. W. Alpha-1-antitrypsin: molecular pathology, leukocytes, and tissue damage. J. Clin. Invest. 78: 1427-1431, 1986. [PubMed: 3537008] [Full Text: https://doi.org/10.1172/JCI112731]

  19. Chappell, S., Daly, L., Morgan, K., Baranes, T. G., Roca, J., Rabinovich, R., Millar, A., Donnelly, S. C., Keatings, V., MacNee, W., Stolk, J., Hiemstra, P., Miniati, M., Monti, S., O'Connor, C. M., Kalsheker, N. Cryptic haplotypes of SERPINA1 confer susceptibility to chronic obstructive pulmonary disease. Hum. Mutat. 27: 103-109, 2006. [PubMed: 16278826] [Full Text: https://doi.org/10.1002/humu.20275]

  20. Chapuis-Cellier, C., Verdier, M., Lepetit, J. C., Fudenberg, H. H., Arnaud, P. Pi-Gm linkage: evidence for linkage in males but not in females and for an effect of the S allele of the Pi system. J. Immunogenet. 8: 257-262, 1981. [PubMed: 6792288] [Full Text: https://doi.org/10.1111/j.1744-313x.1981.tb00767.x]

  21. Clark, P., Breit, S. N., Dawkins, R. L., Penny, R. Genetic study of a family with two members with Weber-Christian disease (panniculitis) and alpha-1-antitrypsin deficiency. Am. J. Med. Genet. 13: 57-62, 1982. [PubMed: 6982619] [Full Text: https://doi.org/10.1002/ajmg.1320130110]

  22. Clark, P., Martin, N. G. An excess of the Pi(s) allele in dizygotic twins and their mothers. Hum. Genet. 61: 171-174, 1982. [PubMed: 6982218] [Full Text: https://doi.org/10.1007/BF00274213]

  23. Cox, D. W., Billingsley, G. D., Mansfield, T. DNA restriction-site polymorphisms associated with the alpha-1-antitrypsin gene. Am. J. Hum. Genet. 41: 891-906, 1987. [PubMed: 2890296]

  24. Cox, D. W., Levison, H. Emphysema of early onset associated with a complete deficiency of alpha-1-antitrypsin (null homozygotes). Am. Rev. Respir. Dis. 137: 371-375, 1988. [PubMed: 3257661] [Full Text: https://doi.org/10.1164/ajrccm/137.2.371]

  25. Cox, D. W., Markovic, V. D., Teshima, I. E. Genes for immunoglobulin heavy chains and for alpha-1-antitrypsin are localized to specific regions of chromosome 14q. Nature 297: 428-430, 1982. [PubMed: 6804874] [Full Text: https://doi.org/10.1038/297428a0]

  26. Cox, D. W., Smyth, S. Risk for liver disease in adults with alpha-1-antitrypsin deficiency. Am. J. Med. 74: 221-227, 1983. [PubMed: 6600583] [Full Text: https://doi.org/10.1016/0002-9343(83)90615-0]

  27. Cox, D. W., Woo, S. L. C., Mansfield, T. DNA restriction fragments associated with alpha-1-antitrypsin indicate a single origin for deficiency allele PI Z. Nature 316: 79-81, 1985. [PubMed: 2989709] [Full Text: https://doi.org/10.1038/316079a0]

  28. Cox, D. W. The effect of neuraminidase on genetic variants of alpha-1-antitrypsin. Am. J. Hum. Genet. 27: 165-177, 1975. [PubMed: 1079112]

  29. Cox, D. W. Genetic variation in alpha-1-antitrypsin. (Editorial) Am. J. Hum. Genet. 30: 660-662, 1978. [PubMed: 311585]

  30. Cox, D. W. Transmission of Z allele from heterozygotes for alpha-1-antitrypsin deficiency. (Letter) Am. J. Hum. Genet. 32: 455-457, 1980. [PubMed: 17948551]

  31. Cox, D. W. New variants of alpha-1-antitrypsin: comparison of Pi typing techniques. Am. J. Hum. Genet. 33: 354-365, 1981. [PubMed: 6972696]

  32. Cox, D. W. Alpha-1-antitrypsin deficiency. In: Scriver, C. R.; Beaudet, A. L.; Sly, W. S.; Valle, D. (eds.): The Metabolic Basis of Inherited Disease. (6th ed.) New York: McGraw-Hill (pub.) 1989.

  33. Cox, D. W. Alpha-1-antitrypsin. In: Scriver, C. R.; Beaudet, A. L.; Sly, W. S.; Valle, D. (eds.): The Metabolic and Molecular Bases of Inherited Disease. Vol. IV. (8th ed.) New York: McGraw-Hill 2001. Pp. 5559-5584.

  34. Crystal, R. G. The alpha-1-antitrypsin gene and its deficiency states. Trends Genet. 5: 411-417, 1989. [PubMed: 2696185] [Full Text: https://doi.org/10.1016/0168-9525(89)90200-x]

  35. Crystal, R. G. Alpha-1-antitrypsin deficiency, emphysema, and liver disease: genetic basis and strategies for therapy. J. Clin. Invest. 85: 1343-1352, 1990. [PubMed: 2185272] [Full Text: https://doi.org/10.1172/JCI114578]

  36. Curiel, D., Brantly, M., Curiel, E., Stier, L., Crystal, R. G. Alpha-1-antitrypsin deficiency caused by the alpha-1-antitrypsin null(Mattawa) gene: an insertion mutation rendering the alpha-1-antitrypsin gene incapable of producing alpha-1-antitrypsin. J. Clin. Invest. 83: 1144-1152, 1989. Note: Erratum: J. Clin. Invest. 84: following 1041, 1989. [PubMed: 2539391] [Full Text: https://doi.org/10.1172/JCI113994]

  37. Curiel, D. T., Chytil, A., Courtney, M., Crystal, R. G. Serum alpha-1-antitrypsin deficiency associated with the common S-type (glu264-to-val) mutation results from intracellular degradation of alpha-1-antitrypsin prior to secretion. J. Biol. Chem. 264: 10477-10486, 1989. [PubMed: 2567291]

  38. Curiel, D. T., Holmes, M. D., Okayama, H., Brantly, M. L., Vogelmeier, C., Travis, W. D. Stier, L. E.; Perks, W. H. and Crystal, R. G.: Molecular basis of the liver and lung disease associated with the alpha-1-antitrypsin deficiency allele M(Malton). J. Biol. Chem. 264: 13938-13945, 1989. [PubMed: 2788166]

  39. Curiel, D. T., Vogelmeier, C., Hubbard, R. C., Stier, L. E., Crystal, R. G. Molecular basis of alpha-1-antitrypsin deficiency and emphysema associated with the alpha-1-antitrypsin M(Mineral Springs) allele. Molec. Cell. Biol. 10: 47-56, 1990. [PubMed: 1967187] [Full Text: https://doi.org/10.1128/mcb.10.1.47-56.1990]

  40. Darlington, G. J., Astrin, K. H., Muirhead, S. P., Desnick, R. J., Smith, M. Assignment of human alpha-1-antitrypsin to chromosome 14 by somatic cell hybrid analysis. Proc. Nat. Acad. Sci. 79: 870-873, 1982. [PubMed: 6801664] [Full Text: https://doi.org/10.1073/pnas.79.3.870]

  41. DeCroo, S., Kamboh, M. I., Ferrell, R. E. Population genetics of alpha-1-antitrypsin polymorphism in US whites, US blacks and African blacks. Hum. Hered. 41: 215-221, 1991. [PubMed: 1783408] [Full Text: https://doi.org/10.1159/000154004]

  42. Dycaico, M. J., Grant, S. G. N., Felts, K., Nichols, W. S., Geller, S. A., Hager, J. H., Pollard, A. J., Kohler, S. W., Short, H. P., Jirik, F. R., Hanahan, D., Sorge, J. A. Neonatal hepatitis induced by alpha-1-antitrypsin: a transgenic mouse model. Science 242: 1409-1415, 1988. [PubMed: 3264419] [Full Text: https://doi.org/10.1126/science.3264419]

  43. Emmerich, J., Alhenc-Gelas, M., Gandrille, S., Guichet, C., Fiessinger, J.-N., Aiach, M. Mechanism of protein C deficiency in a patient with arginine 358 alpha(1)-antitrypsin (Pittsburgh mutation): role in the maintenance of hemostatic balance. J. Lab. Clin. Med. 125: 531-539, 1995. [PubMed: 7706910]

  44. Faber, J. P., Poller, W., Weidinger, S., Kirchgesser, M., Schwaab, R., Bidlingmaier, S., Olek, K. Identification and DNA sequence analysis of 15 new alpha-1-antitrypsin variants, including two Pi* Q alleles and one deficient Pi* M allele. Am. J. Hum. Genet. 55: 1113-1121, 1994. [PubMed: 7977369]

  45. Faber, J. P., Weidinger, S., Goedde, H.-W., Olek, K. The deficient alpha-1-antitrypsin phenotype Pi P is associated with an A-to-T transversion in exon III of the gene. (Letter) Am. J. Hum. Genet. 45: 161-163, 1989. [PubMed: 2787118]

  46. Faber, J. P., Weidinger, S., Olek, K. Sequence data of two rare deficient alpha-1-antitrypsin phenotypes. (Abstract) Cytogenet. Cell Genet. 51: 995-996, 1989.

  47. Faber, J.-P., Weidinger, S., Olek, K. Sequence data of the rare deficient alpha-1-antitrypsin variant PI Z(Augsburg). Am. J. Hum. Genet. 46: 1158-1162, 1990. [PubMed: 2339709]

  48. Fagerhol, M. K., Braend, M. Serum prealbumin: polymorphism in man. Science 149: 986-987, 1965. [PubMed: 5827347] [Full Text: https://doi.org/10.1126/science.149.3687.986]

  49. Fagerhol, M. K., Cox, D. W. The Pi polymorphism: genetic, biochemical, and clinical aspects of human alpha-1-antitrypsin. Adv. Hum. Genet. 11: 1-62, 1981. [PubMed: 6168185]

  50. Fagerhol, M. K., Gedde-Dahl, T., Jr. Genetics of the Pi serum types: family studies of the inherited variants of serum alpha-1-antitrypsin. Hum. Hered. 19: 354-359, 1969. [PubMed: 5366282] [Full Text: https://doi.org/10.1159/000152238]

  51. Fagerhol, M. K., Hauge, H. E. The Pi phenotype MP: discovery of a ninth allele belonging to the system of inherited variants of serum alpha-1-antitrypsin. Vox Sang. 15: 396-400, 1968. [PubMed: 5699691] [Full Text: https://doi.org/10.1111/j.1423-0410.1968.tb04081.x]

  52. Fagerhol, M. K., Laurell, C. B. The polymorphism of 'prealbumins' and alpha-1-antitrypsin in human sera. Clin. Chim. Acta 16: 199-203, 1967. [PubMed: 4166396] [Full Text: https://doi.org/10.1016/0009-8981(67)90181-7]

  53. Fagerhol, M. K., Laurell, C. B. The Pi system--inherited variants of serum alpha-1-antitrypsin. Prog. Med. Genet. 7: 96-111, 1970. [PubMed: 4911922]

  54. Fagerhol, M. K., Tenfjord, O. W. Serum Pi types in some European, American, Asian and African populations. Acta Path. Microbiol. Scand. 72: 601-608, 1968. [PubMed: 5681806] [Full Text: https://doi.org/10.1111/j.1699-0463.1968.tb00472.x]

  55. Fagerhol, M. K. The Pi system: genetic variants of serum alpha-1-antitrypsin. Ser. Haemat. 1: 153-161, 1968.

  56. Fraizer, G. C., Harrold, T. R., Hofker, M. H., Cox, D. W. In-frame single codon deletion in the M(Malton) deficiency allele of alpha-1-antitrypsin. Am. J. Hum. Genet. 44: 894-902, 1989. [PubMed: 2786335]

  57. Fraizer, G. C., Siewertsen, M. A., Hofker, M. H., Brubacher, M. G., Cox, D. W. A null deficiency allele of alpha-1-antitrypsin, Q0-Ludwigshafen, with altered tertiary structure. J. Clin. Invest. 86: 1878-1884, 1990. [PubMed: 2254451] [Full Text: https://doi.org/10.1172/JCI114919]

  58. Fraizer, G. C., Siewertsen, M., Harold, T. R., Cox, D. W. Deletion/frameshift mutation in the alpha-1-antitrypsin null allele, PI*Q0(Bolton). Hum. Genet. 83: 377-382, 1989. [PubMed: 2807278] [Full Text: https://doi.org/10.1007/BF00291385]

  59. Frants, R., Eriksson, A. W. A new unstable Pi M variant of alpha-1-antitrypsin in a Finnish isolate. Hum. Hered. 30: 333-342, 1980. [PubMed: 6971248] [Full Text: https://doi.org/10.1159/000153154]

  60. Garver, R. I., Jr., Mornex, J.-F., Nukiwa, T., Brantly, M., Courtney, M., LeCocq, J.-P., Crystal, R. G. Alpha-1-antitrypsin deficiency and emphysema caused by homozygous inheritance of non-expressing alpha-1-antitrypsin genes. New Eng. J. Med. 314: 762-766, 1986. [PubMed: 3485249] [Full Text: https://doi.org/10.1056/NEJM198603203141207]

  61. Gedde-Dahl, T., Jr., Cook, P. J. L., Fagerhol, M. K., Pierce, J. A. Improved estimate of the Gm-Pi linkage. Ann. Hum. Genet. 39: 43-50, 1975. [PubMed: 810069] [Full Text: https://doi.org/10.1111/j.1469-1809.1975.tb00106.x]

  62. Gedde-Dahl, T., Jr., Cook, P. J. L., Fagerhol, M. K., Pierce, J. A. The Gm-Pi linkage: a summary estimate. Birth Defects Orig. Art. Ser. XI(3): 157-158, 1975. Note: Alternate: Cytogenet. Cell Genet. 14: 327-328, 1975.

  63. Gedde-Dahl, T., Jr., Fagerhol, M. K., Cook, P. J. L., Noades, J. Autosomal linkage between the Gm and Pi loci in man. Ann. Hum. Genet. 35: 393-400, 1972. [PubMed: 5073686]

  64. Gedde-Dahl, T., Jr., Frants, R., Olaisen, B., Eriksson, A. W., van Loghem, E., Lamm, L. The Gm-Pi linkage heterogeneity in view of Pi M subtypes. Ann. Hum. Genet. 45: 143-153, 1981. [PubMed: 6797346] [Full Text: https://doi.org/10.1111/j.1469-1809.1981.tb00316.x]

  65. Graham, A., Hayes, K., Weidinger, S., Newton, C. R., Markham, A. F., Kalsheker, N. A. Characterisation of the alpha-1-antitrypsin M3 gene, a normal variant. Hum. Genet. 85: 381-382, 1990. [PubMed: 2394452] [Full Text: https://doi.org/10.1007/BF00206766]

  66. Graham, A., Kalsheker, N. A., Bamforth, F. J., Newton, C. R., Markham, A. F. Molecular characterization of two alpha-1-antitrypsin deficiency variants: proteinase inhibitor (Pi) null(Newport) (gly(115)-to-ser) and (Pi) Z Wrexham (ser(-19)-to-leu). Hum. Genet. 85: 537-540, 1990. [PubMed: 2227940] [Full Text: https://doi.org/10.1007/BF00194233]

  67. Graham, A., Kalsheker, N. A., Newton, C. R., Bamforth, F. J., Powell, S. J., Markham, A. F. Molecular characterisation of three alpha-1-antitrypsin deficiency variants: proteinase inhibitor (Pi) Null (Cardiff) (asp256-to-val), Pi M(Malton) (phe51-deletion) and Pi I (arg39-to-cys). Hum. Genet. 84: 55-58, 1989. [PubMed: 2606478] [Full Text: https://doi.org/10.1007/BF00210671]

  68. Green, C., Brown, G., Dafforn, T. R., Reichhart, J.-M., Morley, T., Lomas, D. A., Gubb, D. Drosophila necrotic mutations mirror disease-associated variants of human serpins. Development 130: 1473-1478, 2003. [PubMed: 12588861] [Full Text: https://doi.org/10.1242/dev.00350]

  69. Hafeez, W., Ciliberto, G., Perlmutter, D. H. Constitutive and modulated expression of the human alpha-1 antitrypsin gene: different transcriptional initiation sites used in three different cell types. J. Clin. Invest. 89: 1214-1222, 1992. [PubMed: 1556183] [Full Text: https://doi.org/10.1172/JCI115705]

  70. Hidvegi, T., Ewing, M., Hale, P., Dippold, C., Beckett, C., Kemp, C., Maurice, N., Mukherjee, A., Goldbach, C., Watkins, S., Michalopoulos, G., Perlmutter, D. H. An autophagy-enhancing drug promotes degradation of mutant alpha-1-antitrypsin Z and reduces hepatic fibrosis. Science 329: 229-232, 2010. [PubMed: 20522742] [Full Text: https://doi.org/10.1126/science.1190354]

  71. Hildesheim, J., Kinsley, G., Bissell, M., Pierce, J., Brantly, M. Genetic diversity from a limited repertoire of mutations on different common allelic backgrounds: alpha-1-antitrypsin deficiency variant P(Duarte). Hum. Mutat. 2: 221-228, 1993. [PubMed: 8364590] [Full Text: https://doi.org/10.1002/humu.1380020311]

  72. Hofker, M. H., Nelen, M., Klasen, E. C., Nukiwa, T., Curiel, D., Crystal, R. G., Frants, R. R. Cloning and characterization of an alpha-1-antitrypsin like gene 12 kb downstream of the genuine alpha-1-antitrypsin gene. Biochem. Biophys. Res. Commun. 155: 634-642, 1988. [PubMed: 2901833] [Full Text: https://doi.org/10.1016/s0006-291x(88)80542-4]

  73. Hofker, M. H., Nukiwa, T., van Paassen, H. M. B., Nelen, M., Kramps, J. A., Klasen, E. C., Frants, R. R., Crystal, R. G. A pro-to-leu substitution in codon 369 of the alpha-1-antitrypsin deficiency variant PI MHeerlen. Hum. Genet. 81: 264-268, 1989. [PubMed: 2784123] [Full Text: https://doi.org/10.1007/BF00279001]

  74. Holmes, M. D., Brantly, M. L., Crystal, R. G. Molecular analysis of the heterogeneity among the P-family of alpha-1-antitrypsin alleles. Am. Rev. Resp. Dis. 142: 1185-1192, 1990. [PubMed: 2240842] [Full Text: https://doi.org/10.1164/ajrccm/142.5.1185]

  75. Holmes, M. D., Brantly, M. L., Curiel, D. T., Weidinger, S., Crystal, R. G. Characterization of the normal alpha-1-antitrypsin allele V(Munich): a variant associated with a unique protein isoelectric focusing pattern. Am. J. Hum. Genet. 46: 810-816, 1990. [PubMed: 2316526]

  76. Holmes, M. D., Brantly, M. L., Fells, G. A., Crystal, R. G. Alpha-1-antitrypsin W(Bethesda): molecular basis of an unusual alpha-1-antitrypsin deficiency variant. Biochem. Biophys. Res. Commun. 170: 1013-1020, 1990. [PubMed: 2390072] [Full Text: https://doi.org/10.1016/0006-291x(90)90493-7]

  77. Hug, G., Chuck, G., Fagerhol, M. K. Pi(P-Clifton): a new alpha(1) antitrypsin allele in an American Negro family. J. Med. Genet. 18: 43-45, 1981. [PubMed: 6973024] [Full Text: https://doi.org/10.1136/jmg.18.1.43]

  78. Hug, G., Chuck, G., Slemmer, T. M., Fagerhol, M. K. Pi (Ecincinnati): a new alpha-1-antitrypsin allele in three Negro families. Hum. Genet. 54: 361-364, 1980. [PubMed: 6967446] [Full Text: https://doi.org/10.1007/BF00291583]

  79. Hutchison, D. C. S. Alpha-1-antitrypsin deficiency in Europe: geographical distribution of Pi types S and Z. Respir. Med. 92: 367-377, 1998. [PubMed: 9692092] [Full Text: https://doi.org/10.1016/s0954-6111(98)90278-5]

  80. Iammarino, R. M., Wagener, D. K., Allen, R. C. Segregation distortion of the alpha-1-antitrypsin Pi Z allele. Am. J. Hum. Genet. 31: 508-517, 1979. [PubMed: 314754]

  81. Kalsheker, N. A., Hodgson, I. J., Watkins, G. L., White, J. P., Morrison, H. M., Stockley, R. A. Deoxyribonucleic acid (DNA) polymorphism of the alpha-1-antitrypsin gene in chronic lung disease. Brit. Med. J. 294: 1511-1514, 1987. [PubMed: 3038256] [Full Text: https://doi.org/10.1136/bmj.294.6586.1511]

  82. Kalsheker, N., Hayes, K., Weidinger, S., Graham, A. What is Pi (proteinase inhibitor) null or PiQ0? A problem highlighted by the alpha-1-antitrypsin M(Heerlen) mutation. J. Med. Genet. 29: 27-29, 1992. [PubMed: 1552539] [Full Text: https://doi.org/10.1136/jmg.29.1.27]

  83. Kamimoto, T., Shoji, S., Hidvegi, T., Mizushima, N., Umebayashi, K., Perlmutter, D. H., Yoshimori, T. Intracellular inclusions containing mutant alpha(1)-antitrypsin Z are propagated in the absence of autophagic activity. J. Biol. Chem. 281: 4467-4476, 2006. [PubMed: 16365039] [Full Text: https://doi.org/10.1074/jbc.M509409200]

  84. Kawakami, Y., Irie, T., Kishi, F., Asanuma, Y., Shida, A., Yoshikawa, T., Kamishima, K., Hasegawa, H., Murao, M. Familial aggregation of abnormal ventilatory control and pulmonary function in chronic obstructive pulmonary disease. Europ. J. Resp. Dis. 62: 56-64, 1981. [PubMed: 7227484]

  85. Kelly, E., Greene, C. M., Carroll, T. P., McElvaney, N. G., O'Neill, S. J. Selenoprotein S/SEPS1 modifies endoplasmic reticulum stress in Z variant alpha-1-antitrypsin deficiency. J. Biol. Chem. 284: 16891-16897, 2009. [PubMed: 19398551] [Full Text: https://doi.org/10.1074/jbc.M109.006288]

  86. Keyeux, G., Gilgenkrantz, S., Lefranc, G., Lefranc, M.-P. Molecular characterization of a ring chromosome 14 showing that the PI locus is centromeric to the D14S1 and IGH loci. Hum. Genet. 82: 219-222, 1989. [PubMed: 2543620] [Full Text: https://doi.org/10.1007/BF00291158]

  87. Kramps, J. A., Brouwers, J. W., Maesen, F., Dijkman, J. H. Pi(M-Heerlen), a Pi(M) allele resulting in very low alpha-1-antitrypsin serum levels. Hum. Genet. 59: 104-107, 1981. [PubMed: 6976926] [Full Text: https://doi.org/10.1007/BF00293055]

  88. Kueppers, F., Bearn, A. G. An inherited alpha-1-antitrypsin variant. Humangenetik 4: 217-220, 1967. [PubMed: 6080804] [Full Text: https://doi.org/10.1007/BF00292195]

  89. Kueppers, F., Christopherson, M. J. Alpha-1-antitrypsin: further genetic heterogeneity revealed by isoelectric focusing. Am. J. Hum. Genet. 30: 359-365, 1978. [PubMed: 309724]

  90. Kurachi, K., Chandra, T., Degen, S. J. F., White, T. T., Marchioro, T. L., Woo, S. L. C., Davie, E. W. Cloning and sequence of cDNA coding for alpha-1-antitrypsin. Proc. Nat. Acad. Sci. 78: 6826-6830, 1981. [PubMed: 7031661] [Full Text: https://doi.org/10.1073/pnas.78.11.6826]

  91. Lace, B., Sveger, T., Krams, A., Cernevska, G., Krumina, A. Age of SERPINA1 gene PI Z mutation: Swedish and Latvian population analysis. Ann. Hum. Genet. 72: 300-304, 2008. [PubMed: 18294358] [Full Text: https://doi.org/10.1111/j.1469-1809.2008.00431.x]

  92. Lai, E. C., Kao, F.-T., Law, M. L., Woo, S. L. C. Assignment of the alpha-1-antitrypsin gene and a sequence-related gene to human chromosome 14 by molecular hybridization. Am. J. Hum. Genet. 35: 385-392, 1983. [PubMed: 6602546]

  93. Laubach, V. E., Ryan, W. J., Brantly, M. Characterization of a human alpha-1-antitrypsin null allele involving aberrant mRNA splicing. Hum. Molec. Genet. 2: 1001-1005, 1993. [PubMed: 8364536] [Full Text: https://doi.org/10.1093/hmg/2.7.1001]

  94. Ledbetter, S. A., Ledbetter, D. H., Ledley, F. D., Woo, S. Localization of phenylalanine hydroxylase (PAH) and alpha-1 antitrypsin (AAT) loci in mouse genome by synteny and in situ hybridization. (Abstract) Am. J. Hum. Genet. 41: A173, 1987.

  95. Lewis, J. H., Iammarino, R. M., Spero, J. A., Hasiba, U. Antithrombin Pittsburgh: an alpha-1-antitrypsin variant causing hemorrhagic disease. Blood 51: 129-137, 1978. [PubMed: 412531]

  96. Lieberman, J., Borhani, N. O., Feinleib, M. Alpha-1-antitrypsin deficiency in twins and parents-of-twins. Clin. Genet. 15: 29-36, 1979. [PubMed: 310370] [Full Text: https://doi.org/10.1111/j.1399-0004.1979.tb02026.x]

  97. Long, G. L., Chandra, T., Woo, S. L. C., Davie, E. W., Kurachi, K. Complete sequence of the cDNA for human alpha-1-antitrypsin and the gene for the S variant. Biochemistry 23: 4828-4837, 1984. [PubMed: 6093867] [Full Text: https://doi.org/10.1021/bi00316a003]

  98. Lopez, V., Oetliker, O., Colombo, J. P., Butler, R. Ein Fall von familiaerem alpha-1-Antitrypsinmangel. Helv. Paediat. Acta 19: 296-303, 1964. [PubMed: 14229909]

  99. Lovegrove, J. U., Jeremiah, S., Gillett, G. T., Temple, I. K., Povey, S., Whitehouse, D. B. A new alpha 1-antitrypsin mutation, thr-met 85, (PI Z-Bristol) associated with novel electrophoresis properties. Ann. Hum. Genet. 61: 385-391, 1997. [PubMed: 9459000] [Full Text: https://doi.org/10.1046/j.1469-1809.1997.6150385.x]

  100. Martorana, P. A., Brand, T., Gardi, C., van Even, P., de Santi, M. M., Calzoni, P., Marcolongo, P., Lungarella, G. The pallid mouse: a model of genetic alpha-1-antitrypsin deficiency. Lab. Invest. 68: 233-241, 1993. [PubMed: 8441253]

  101. Matsunaga, E., Shiokawa, S., Nakamura, H., Maruyama, T., Tsuda, K., Fukumaki, Y. Molecular analysis of the gene of the alpha-1-antitrypsin deficiency variant, M(Nichinan). Am. J. Hum. Genet. 46: 602-612, 1990. [PubMed: 2309708]

  102. Morgan, K., Scobie, G., Kalsheker, N. A. Point mutation in a 3-prime flanking sequence of the alpha-1-antitrypsin gene associated with chronic respiratory disease occurs in a regulatory sequence. Hum. Molec. Genet. 2: 253-257, 1993. [PubMed: 8499914] [Full Text: https://doi.org/10.1093/hmg/2.3.253]

  103. Munch, J., Standker, L., Adermann, K., Schulz, A., Schindler, M., Chinnadurai, R., Pohlmann, S., Chaipan, C., Biet, T., Peters, T., Meyer, B., Wilhelm, D., Lu, H., Jing, W., Jiang, S., Forssmann, W.-G., Kirchhoff, F. Discovery and optimization of a natural HIV-1 entry inhibitor targeting the gp41 fusion peptide. Cell 129: 263-275, 2007. [PubMed: 17448989] [Full Text: https://doi.org/10.1016/j.cell.2007.02.042]

  104. Nakamura, H., Ogawa, A., Hisano, S., Fukuma, M., Tachibana, N., Tsuda, K. A family with a new deficient variant of alpha-1-antitrypsin PiM(Nichinan): with special reference to diastase-resistant, periodic acid-Schiff positive globules in the liver cells. J. Jpn. Soc. Intern. Med. 69: 47-54, 1980.

  105. Nukiwa, T., Brantly, M., Garver, R., Paul, L., Courtney, M., LeCocq, J.-P., Crystal, R. G. Evaluation of 'at risk' alpha-1-antitrypsin genotype SZ with synthetic oligonucleotide gene probes. J. Clin. Invest. 77: 528-537, 1986. [PubMed: 3484754] [Full Text: https://doi.org/10.1172/JCI112333]

  106. Nukiwa, T., Brantly, M. L., Ogushi, F., Fells, G. A., Crystal, R. G. Characterization of the gene and protein of the common alpha-1-antitrypsin normal M2 allele. Am. J. Hum. Genet. 43: 322-330, 1988. [PubMed: 2901226]

  107. Nukiwa, T., Satoh, K., Brantly, M. L., Ogushi, F., Fells, G. A., Courtney, M., Crystal, R. G. Identification of a second mutation in the protein-coding sequence of the Z type alpha-1-antitrypsin gene. J. Biol. Chem. 261: 15989-15994, 1986. Note: Erratum: J. Biol. Chem. 262: 10412 only, 1987. [PubMed: 3491072]

  108. Nukiwa, T., Takahashi, H., Brantly, M., Courtney, M., Crystal, R. G. Alpha-1-antitrypsin null (Granite Falls), a nonexpressing alpha-1-antitrypsin gene associated with a frameshift to stop mutation in a coding exon. J. Biol. Chem. 262: 11999-12004, 1987. [PubMed: 3040726]

  109. Okayama, H., Brantly, M., Holmes, M., Crystal, R. G. Characterization of the molecular basis of the alpha-1-antitrypsin F allele. Am. J. Hum. Genet. 48: 1154-1158, 1991. [PubMed: 2035534]

  110. Owen, M. C., Brennan, S. O., Lewis, J. H., Carrell, R. W. Mutation of antitrypsin to antithrombin: alpha-1-antitrypsin Pittsburgh (358 met-to-arg), a fatal bleeding disorder. New Eng. J. Med. 309: 694-698, 1983. [PubMed: 6604220] [Full Text: https://doi.org/10.1056/NEJM198309223091203]

  111. Owen, M. C., Carrell, R. W., Brennan, S. O. The abnormality of the S variant of human alpha-1-antitrypsin. Biochim. Biophys. Acta 453: 257-261, 1976. [PubMed: 1087161] [Full Text: https://doi.org/10.1016/0005-2795(76)90271-3]

  112. Owen, M. C., Carrell, R. W. Alpha-1-antitrypsin: molecular abnormality of S variant. Brit. Med. J. 1: 130-131, 1976. [PubMed: 1082356] [Full Text: https://doi.org/10.1136/bmj.1.6002.130-a]

  113. Pearson, S., Tetri, P., George, D. L., Francke, U. Alpha-1-antitrypsin (PI) expression in rat hepatoma-human somatic cell hybrids: evidence for PI locus on chromosome 14 and for regulatory locus on the X chromosome. (Abstract) Am. J. Hum. Genet. 33: 148A, 1981.

  114. Perlino, E., Cortese, R., Ciliberto, G. The human alpha-1-antitrypsin gene is transcribed from two different promoters in macrophages and hepatocytes. EMBO J. 6: 2767-2771, 1987. [PubMed: 3500042] [Full Text: https://doi.org/10.1002/j.1460-2075.1987.tb02571.x]

  115. Poller, W., Faber, J.-P., Weidinger, S., Olek, K. DNA polymorphisms associated with a new alpha-1-antitrypsin PI Q0 variant (PI Q0-Riedenburg). Hum. Genet. 86: 522-524, 1991. [PubMed: 1673114] [Full Text: https://doi.org/10.1007/BF00194647]

  116. Poller, W., Meisen, C., Olek, K. DNA polymorphisms of the alpha-1-antitrypsin gene region in patients with chronic obstructive pulmonary disease. Europ. J. Clin. Invest. 20: 1-7, 1990. [PubMed: 1969347] [Full Text: https://doi.org/10.1111/j.1365-2362.1990.tb01769.x]

  117. Roychoudhury, A. K., Nei, M. Human Polymorphic Genes: World Distribution. New York: Oxford Univ. Press (pub.) 1988. Pp. 132-135.

  118. Satoh, K., Nukiwa, T., Brantly, M., Garver, R. I., Jr., Hofker, M., Courtney, M., Crystal, R. G. Emphysema associated with complete absence of alpha-1-antitrypsin in serum and the homozygous inheritance of a stop codon in an alpha-1-antitrypsin-coding exon. Am. J. Hum. Genet. 42: 77-83, 1988. Note: Erratum: Am. J. Hum. Genet. 42: 789 only, 1988. [PubMed: 3257351]

  119. Schapira, M., Ramus, M.-A., Jallat, S., Carvallo, D., Courtney, M. Recombinant alpha-1-antitrypsin Pittsburgh (met-358 to arg) is a potent inhibitor of plasma kallikrein and activated factor XII fragment. J. Clin. Invest. 77: 635-637, 1986. [PubMed: 3484756] [Full Text: https://doi.org/10.1172/JCI112347]

  120. Schroeder, W. T., Miller, M. F., Woo, S. L. C., Saunders, G. F. Chromosomal localization of the human alpha-antitrypsin gene (PI) to 14q31-32. Am. J. Hum. Genet. 37: 868-872, 1985. [PubMed: 3876766]

  121. Scott, C. F., Carrell, R. W., Glaser, C. B., Kueppers, F., Lewis, J. H., Colman, R. W. Alpha-1-antitrypsin-Pittsburgh: a potent inhibitor of human plasma factor XIa, kallikrein, and factor XII. J. Clin. Invest. 77: 631-634, 1986. [PubMed: 3484755] [Full Text: https://doi.org/10.1172/JCI112346]

  122. Sefton, L., Kearney, P., Kelsey, G., Povey, S., Wolfe, J. Physical linkage of the genes PI and AACT. (Abstract) Cytogenet. Cell Genet. 51: 1076, 1989.

  123. Seixas, S., Mendonca, C., Costa, F., Rocha, J. Alpha-1-antitrypsin null alleles: evidence for the recurrence of the L353fsX376 mutation and a novel G-to-A transition in position +1 of intron IC affecting normal mRNA splicing. Clin. Genet. 62: 175-180, 2002. [PubMed: 12220457] [Full Text: https://doi.org/10.1034/j.1399-0004.2002.620212.x]

  124. Seri, M., Magi, B., Cellesi, C., Olia, P. M., Renieri, A., De Marchi, M. Molecular characterization of the P and I variants of alpha-1-antitrypsin. Int. J. Clin. Lab. Res. 22: 119-121, 1992. [PubMed: 1504305] [Full Text: https://doi.org/10.1007/BF02591409]

  125. Shapiro, L., Pott, G. B., Ralston, A. H. Alpha-1-antitrypsin inhibits human immunodeficiency virus type 1. FASEB J. 15: 115-122, 2001. [PubMed: 11149899] [Full Text: https://doi.org/10.1096/fj.00-0311com]

  126. Sifers, R. N., Brashears-Macatee, S., Kidd, V. J., Muensch, H., Woo, S. L. C. A frameshift mutation results in a truncated alpha-1-antitrypsin that is retained within the rough endoplasmic reticulum. J. Biol. Chem. 263: 7330-7335, 1988. [PubMed: 3259232]

  127. Soutoglou, E., Talianidis, I. Coordination of PIC assembly and chromatin remodeling during differentiation-induced gene activation. Science 295: 1901-1904, 2002. [PubMed: 11884757] [Full Text: https://doi.org/10.1126/science.1068356]

  128. Takahashi, H., Crystal, R. G. Alpha-1-antitrypsin Null(Isola di Procida): an alpha-1-antitrypsin deficiency allele caused by deletion of all alpha-1-antitrypsin coding exons. Am. J. Hum. Genet. 47: 403-413, 1990. [PubMed: 1975477]

  129. Takahashi, H., Nukiwa, T., Satoh, K., Ogushi, F., Brantly, M., Fells, G., Stier, L., Courtney, M., Crystal, R. G. Characterization of the gene and protein of the alpha-1-antitrypsin 'deficiency' allele M(procida). J. Biol. Chem. 263: 15528-15534, 1988. [PubMed: 3262617]

  130. Turleau, C., de Grouchy, J., Chavin-Colin, F., Dore, F., Seger, J., Dautzenberg, M., Arthuis, M., Jeanson, C. Two patients with interstitial del(14q), one with features of Holt-Oram syndrome: exclusion mapping of PI (alpha-1-antitrypsin). Ann. Genet. 27: 237-240, 1984. [PubMed: 6335371]

  131. van Pel, M., van Os, R., Velders, G. A., Hagoort, H., Heegaard, P. M. H., Lindley, I. J. D., Willemze, R., Fibbe, W. E. Serpina1 is a potent inhibitor of IL-8-induced hematopoietic stem cell mobilization. Proc. Nat. Acad. Sci. 103: 1469-1474, 2006. [PubMed: 16432201] [Full Text: https://doi.org/10.1073/pnas.0510192103]

  132. Vidaud, D., Emmerich, J., Alhenc-Gelas, M., Yvart, J., Fiessinger, J. N., Aiach, M. Met358-to-arg mutation of alpha-1-antitrypsin associated with protein C deficiency in a patient with mild bleeding tendency. J. Clin. Invest. 89: 1537-1543, 1992. [PubMed: 1569192] [Full Text: https://doi.org/10.1172/JCI115746]

  133. Weidinger, S., Jahn, W., Cujnik, F., Schwarzfischer, F. Alpha-1-antitrypsin: evidence for a fifth PI M subtype and a new deficiency allele PI*Z(Augsburg). Hum. Genet. 71: 27-29, 1985. [PubMed: 3875547] [Full Text: https://doi.org/10.1007/BF00295662]

  134. Weitkamp, L. R., Cox, D., Guttormsen, S., Johnston, E., Hempfling, S. Allelic specific heterogeneity in the Pi-Gm linkage group. Cytogenet. Cell Genet. 22: 647-650, 1978. [PubMed: 313312] [Full Text: https://doi.org/10.1159/000131044]

  135. Welch, S. G., McGregor, I. A., Williams, K. Alpha-1-antitrypsin (Pi) phenotypes in a village population from the Gambia, West Africa. Hum. Genet. 53: 233-235, 1980. [PubMed: 6965658] [Full Text: https://doi.org/10.1007/BF00273503]

  136. Whitehouse, D. B., Abbott, C. M., Lovegrove, J. U., McIntosh, I., McMahon, C. J., Mieli-Vergani, G., Mowat, A. P., Hopkinson, D. A. Genetic studies on a new deficiency gene (PI*Z-Tun) at the PI locus. J. Med. Genet. 26: 744-749, 1989. [PubMed: 2575668] [Full Text: https://doi.org/10.1136/jmg.26.12.744]

  137. Wilkie, A. O. M. The molecular basis of genetic dominance. J. Med. Genet. 31: 89-98, 1994. [PubMed: 8182727] [Full Text: https://doi.org/10.1136/jmg.31.2.89]

  138. Yamamoto, Y., Sawa, R., Okamoto, N., Matsui, A., Yanagisawa, M., Ikemoto, S. Deletion 14q(q24.3 to q32.1) syndrome: significance of peculiar facial appearance in its diagnosis, and deletion mapping of Pi (alpha-1-antitrypsin). Hum. Genet. 74: 190-192, 1986. [PubMed: 3490426] [Full Text: https://doi.org/10.1007/BF00282092]

  139. Yoshida, A., Chillar, R., Taylor, J. C. An alpha-1-antitrypsin variant, PiB Alhambra (lys-to-asp, glu-to-asp), with rapid anodal electrophoretic mobility. Am. J. Hum. Genet. 31: 555-563, 1979. [PubMed: 315708]

  140. Yoshida, A., Ewing, C., Wessels, M., Lieberman, J., Gaidulis, L. Molecular abnormality of Pi S variant of human alpha-1-antitrypsin. Am. J. Hum. Genet. 29: 233-239, 1977. [PubMed: 301355]

  141. Yoshida, A., Lieberman, J., Gaidulis, L., Ewing, C. Molecular abnormality of human alpha-1-antitrypsin variant (Pi-ZZ) associated with plasma activity deficiency. Proc. Nat. Acad. Sci. 73: 1324-1328, 1976. [PubMed: 1083527] [Full Text: https://doi.org/10.1073/pnas.73.4.1324]

  142. Yoshida, A., Taylor, J. C., Van den Brock, W. G. M. Structural difference between the normal PiM(1) and the common PiM(2) variant of human alpha-1-antitrypsin. Am. J. Hum. Genet. 31: 564-568, 1979. [PubMed: 315709]

  143. Yuasa, I., Sugimoto, Y., Ichinose, M., Matsumoto, Y., Fukumaki, Y., Sasaki, T., Okada, K. PI*S(Iiyama), a deficiency gene of alpha-1-antitrypsin: evidence for the occurrence in western Japan. Jpn. J. Hum. Genet. 38: 185-191, 1993. [PubMed: 8358043] [Full Text: https://doi.org/10.1007/BF01883709]


Contributors:
Bao Lige - updated : 08/31/2022
Patricia A. Hartz - updated : 7/11/2012
Ada Hamosh - updated : 9/7/2011
Ada Hamosh - updated : 8/17/2010
Cassandra L. Kniffin - updated : 1/13/2009
Paul J. Converse - updated : 8/17/2007
Paul J. Converse - updated : 3/24/2006
Victor A. McKusick - updated : 1/20/2006
Paul J. Converse - updated : 3/14/2005
Victor A. McKusick - updated : 12/4/2003
Victor A. McKusick - updated : 7/10/2003
Ada Hamosh - updated : 4/15/2003
Victor A. McKusick - updated : 11/26/2002
Victor A. McKusick - updated : 6/3/2002
Ada Hamosh - updated : 4/2/2002
Victor A. McKusick - updated : 2/12/2002
Victor A. McKusick - updated : 1/10/2002
Victor A. McKusick - updated : 10/21/1999
Victor A. McKusick - updated : 12/10/1998
Victor A. McKusick - updated : 11/5/1998
Victor A. McKusick - updated : 7/10/1998
Cynthia K. Ewing - updated : 10/23/1996

Creation Date:
Victor A. McKusick : 6/4/1986

Edit History:
carol : 09/30/2022
mgross : 08/31/2022
carol : 07/20/2016
carol : 07/18/2016
carol : 7/15/2016
carol : 7/14/2016
carol : 7/9/2016
carol : 7/9/2016
alopez : 8/7/2013
carol : 5/21/2013
terry : 4/4/2013
terry : 3/15/2013
carol : 8/15/2012
terry : 8/8/2012
terry : 7/11/2012
carol : 5/30/2012
terry : 10/10/2011
mgross : 10/3/2011
alopez : 9/8/2011
terry : 9/7/2011
terry : 9/7/2010
terry : 9/7/2010
terry : 9/7/2010
alopez : 8/18/2010
joanna : 8/17/2010
terry : 8/17/2010
carol : 8/13/2010
carol : 9/15/2009
terry : 6/3/2009
wwang : 4/1/2009
wwang : 1/16/2009
ckniffin : 1/13/2009
terry : 1/8/2009
terry : 1/8/2009
terry : 1/7/2009
terry : 1/7/2009
carol : 10/4/2007
mgross : 8/17/2007
mgross : 3/29/2006
terry : 3/24/2006
carol : 2/14/2006
alopez : 1/25/2006
terry : 1/20/2006
carol : 11/28/2005
mgross : 3/14/2005
mgross : 3/14/2005
mgross : 3/17/2004
alopez : 12/10/2003
terry : 12/4/2003
carol : 11/24/2003
carol : 7/21/2003
terry : 7/10/2003
alopez : 4/17/2003
terry : 4/15/2003
cwells : 11/26/2002
terry : 11/20/2002
cwells : 6/26/2002
terry : 6/3/2002
carol : 4/25/2002
alopez : 4/5/2002
terry : 4/2/2002
terry : 3/13/2002
terry : 2/12/2002
cwells : 1/25/2002
cwells : 1/16/2002
terry : 1/10/2002
alopez : 11/23/1999
carol : 10/21/1999
carol : 12/16/1998
terry : 12/10/1998
carol : 11/16/1998
terry : 11/5/1998
dkim : 7/24/1998
dkim : 7/24/1998
carol : 7/15/1998
terry : 7/10/1998
terry : 5/29/1998
terry : 1/29/1998
mark : 10/14/1997
mark : 10/14/1997
dholmes : 8/14/1997
alopez : 7/31/1997
jenny : 7/9/1997
joanna : 7/3/1997
joanna : 6/24/1997
terry : 1/10/1997
mark : 12/29/1996
jamie : 10/23/1996
jamie : 10/16/1996
jamie : 10/14/1996
mark : 3/25/1996
terry : 7/10/1995
mark : 6/13/1995
pfoster : 5/2/1995
davew : 8/18/1994
jason : 6/17/1994
warfield : 4/4/1994