REVIEW

Mutational Spectra of PTEN/MMAC1 Gene: a Tumor Suppressor With Lipid Phosphatase Activity

Iqbal Unnisa Ali, Lynn M. Schriml, Michael Dean

Affiliations of authors: I. U. Ali, Division of Cancer Prevention, National Cancer Institute, Bethesda, MD; L. M. Schriml, M. Dean, Laboratory of Genomic Diversity, National Cancer Institute-Frederick Cancer Research and Development Center, Frederick, MD.

Correspondence to: Iqbal U. Ali, Ph.D., National Institutes of Health, Executive Plaza North, Rm. 201, Bethesda, MD 20892-7332 (e-mail: ia1t{at}nih.gov).


    ABSTRACT
 Top
 Abstract
 Introduction
 I. Germline PTEN/MMAC1 Mutations...
 II. Inactivation of PTEN/MMAC1...
 III. PTEN/MMAC1 as Tumor...
 Notes
 References
 
PTEN/MMAC1 (phosphatase, tensin homologue/mutated in multiple advanced cancers) is a tumor suppressor protein that has sequence homology with dual-specificity phosphatases, which are capable of dephosphorylating both tyrosine phosphate and serine/threonine phosphate residues on proteins. The in vivo function of PTEN/MMAC1 appears to be dephosphorylation of phosphotidylinositol 3,4,5-triphosphate. The PTEN/MMAC1 gene is mutated in the germline of patients with rare autosomal dominant cancer syndromes and in subsets of specific cancers. Here we review the mutational spectra of the PTEN/MMAC1 gene in tumors from various tissues, especially endometrium, brain, prostate, and ovary, in which the gene is inactivated very frequently. Germline and somatic mutations in the PTEN/MMAC1 gene occur mostly in the protein coding region and involve the phosphatase domain and poly(A)6 stretches. Compared with germline alterations found in the PTEN/MMAC1 gene, there is a substantially increased frequency of frameshift mutations in tumors. Glioblastomas and endometrial carcinomas appear to have distinct mutational spectra, probably reflecting differences in the underlying mechanisms of inactivation of the PTEN/MMAC1 gene in the two tissue types. Also, depending on the tissue type, the gene appears to be involved in the initiation or the progression of cancers. Further understanding of PTEN/MMAC1 gene mutations in different tumors and the physiologic consequences of these mutations is likely to open up new therapeutic opportunities for targeting this critical gene.



    INTRODUCTION
 Top
 Abstract
 Introduction
 I. Germline PTEN/MMAC1 Mutations...
 II. Inactivation of PTEN/MMAC1...
 III. PTEN/MMAC1 as Tumor...
 Notes
 References
 
Genetic alterations in a variety of critical genes underlie the formation and progression of human cancers. Some of these genes, such as tumor suppressors p53 (also known as TP53), Rb (retinoblastoma protein), and p16, are involved in central processes, such as transcriptional modulation and cell cycle regulation, and are frequently mutated in human cancers. Inactivation of another tumor suppressor gene PTEN/MMAC1 (phosphatase, tensin homologue/mutated in multiple advanced cancers) has been reported in several types of human tumors, including those from brain, breast, endometrium, kidney, and prostate (1-3). The name PTEN reflects the presence of the protein tyrosine phosphatase domain as well as the domain with homology to tensin. MMAC1 conveys the fact that the gene is mutated in multiple advanced cancers. Considerable information has accumulated within the last year on the mutations in the PTEN/MMAC1 gene in a wide variety of human cancers. In this review, we present the mutational spectra of the PTEN/MMAC1 gene in the germline as well as in tumors of various tissue types. The information was searched in the English language literature by use of the MEDLINE® database. We briefly discuss the enzymatic function of PTEN/MMAC1 with the purpose of understanding possible associations between distinct mutations and functional changes in specific malignancies.

PTEN/MMAC1 is located on chromosome 10q23.3 and encodes a predicted protein product of 403 amino acids containing a protein tyrosine phosphatase domain with a critical motif, IHCKAGKGRTG, found in dual specificity phosphatases (1-3), which dephosphorylate tyrosine as well as serine and threonine residues. The protein also has extensive identity to two cytoskeletal proteins—tensin, which appears to interact with actin filaments in focal adhesions, and auxilin, which is involved in synaptic vesicle transport (1-3). Mapping of homozygous deletions on chromosome 10q23 followed by representational difference analysis and positional cloning led to the isolation of the gene independently by two groups (1,2). Another approach utilized the conserved sequence motifs of phosphatases to isolate a novel protein tyrosine phosphatase, which turned out to be identical with PTEN/MMAC1 and whose transcription was modulated by transforming growth factor-ß (TGF-ß), giving the protein another name, TEP1 (TGF-ß-regulated and epithelial cell-enriched phosphatase) (3).

Three lines of evidence have established PTEN/MMAC1 as a tumor suppressor gene. First, germline mutations in PTEN/MMAC1 are associated with autosomal dominant hamartomatous and often cancer-predisposing syndromes, Cowden's disease and Bannayan-Zonana syndrome (see Section I). Second, the PTEN/MMAC1 gene is homozygously inactivated in a variety of sporadic human cancers (see Section II). Third, it is a highly conserved protein, with phosphatase activity capable of dephosphorylating tyrosine and serine/threonine residues of proteins as well as position 3 on the inositol ring of the lipid phosphatidylinositol 3,4,5-triphosphate (PIP3). Introduction of the PTEN/MMAC1 gene into cancer cell lines leads to growth suppression, and the heterozygous or homozygous inactivation of the gene in the mouse germline results in increased tumor susceptibility or embryonic lethality, respectively (see Section III).


    I. GERMLINE PTEN/MMAC1 MUTATIONS IN PATIENTS WITH HAMARTOMA SYNDROMES AND PREDISPOSITION TO CANCERS
 Top
 Abstract
 Introduction
 I. Germline PTEN/MMAC1 Mutations...
 II. Inactivation of PTEN/MMAC1...
 III. PTEN/MMAC1 as Tumor...
 Notes
 References
 
The hamartomatous polyposis syndromes, Cowden's disease (Mendelian Inheritance in Man [MIM] database No. 158350), Bannayan-Zonana syndrome (MIM No. 153480), and juvenile polyposis coli (MIM No. 174900), are rare autosomal dominant disorders with the common feature of the development of hamartomas, which are hyperplastic, disorganized, and nonmalignant growths. All three syndromes appear to have distinct clinical features as well as partial overlap in their clinical phenotype. The predominant phenotype of Cowden's disease is hamartomas of the skin; other organ sites developing hamartomas include breast, thyroid, endometrium, gastrointestinal tract, and central nervous system, with breast cancers developing in 25%-50% of affected women and thyroid cancer in 3%-10% of affected individuals (4,5). Lhermitte-Duclos disease, which is characterized by an unusual central nervous system tumor, cerebellar dysplastic gangliocytoma, is also associated with Cowden's disease (6). Juvenile polyposis coli patients develop hamartomatous polyps of the digestive tract and are reported to be predisposed to gastrointestinal cancer and possibly to pancreatic cancer (7), whereas in the case of Bannayan-Zonana syndrome, no definite documentation exists regarding an increased risk of malignancy (8).

Germline mutations in the PTEN/MMAC1 gene have been detected in Cowden's disease and in Bannayan-Zonana syndrome families (9-25). In juvenile polyposis coli syndrome, however, the presence of PTEN/MMAC1 germline mutations has been controversial (26,27) for multiple reasons, including overlap of phenotypic features with Cowden's disease and Bannayan-Zonana syndrome, age-related penetrance of the disease, and genetic heterogeneity (28,29). Recently, linkage for juvenile polyposis coli was established on the chromosomal region 18q21.1 (30), and germline mutations in the SMAD4 tumor suppressor gene were detected in a subset of juvenile polyposis coli families (31).

The frequency and types of mutations reported in multiple studies in Cowden's disease and in Bannayan-Zonana syndrome families are shown in Fig. 1.Go Of the 95 germline mutations reported, 25 are frameshift mutations and the rest are point mutations comprising 35 nonsense mutations, 24 missense mutations, and 11 base substitutions generating splice variants. Altogether, 75% of the germline mutations corresponding to nonsense, frameshift, and splice-site mutations result in the generation of a truncated protein. A very high proportion of the missense mutations, 13 (54%) of 24, cluster in the phosphatase core motif beween residues 122 and 132 in exon 5, altering primarily glycine 129 and arginine 130. A similar proportion, 19 (54%) of 35 nonsense mutations, target three codons, 130 and 157 in exon 5 and 233 in exon 7 of the PTEN/MMAC1 gene. Residue 233 is altered by nonsense mutations in one Bannayan-Zonana syndrome and in five Cowden's disease families (9,11,14,15).



View larger version (15K):
[in this window]
[in a new window]
 
Fig. 1. Germline PTEN/MMAC1 mutations in patients with Cowden's disease and Bannayan-Zonana syndrome. Red = frameshift mutations; green = missense mutations; blue = nonsense mutations; black arrows below the bar = splice-site variants generated by point mutations; arrows marked with B = mutations in patients with Bannayan-Zonnana syndrome; and unmarked arrows = mutations in patients with Cowden's disease. Structural and functional domains of the PTEN/MMAC1 gene are shown. Black vertical lines inside the bar = intron/exon boundaries. Exon numbers are shown under the bar. The strongest homology region to tensin/auxilin is in blue, with the hatched region showing the phosphatase domain and the solid red showing the phosphatase core motif. Poly(A)6 stretches between codons 265-267 (in exon 7) and codons 321-323 (in exon 8) and the palindromic sequence preceding the second poly(A)6 stretch encompassing codons 317-320 are shown in orange. Tyrosine phosphorylation sites with tyrosine residues at codons 240 (in exon 7), 315 (in exon 8), and 336 (in exon 8) are shown in purple, and serine phosphorylation sites with serine residues at codons 338 (in exon 8) and 355 (in exon 9) are shown in green.

 
A relatively low frequency or even absence of PTEN/MMAC1 mutations in some Cowden's disease and Bannayan-Zonana syndrome (10,22,32) patients has been reported, suggesting genetic heterogeneity. Nevertheless, the germline mutations reported so far argue that, in most Cowden's disease families and in a subset of Bannayan-Zonana syndrome families, PTEN/MMAC1 is the susceptibility gene and that these hamartomatous polyposis syndromes are allelic to each other. Furthermore, deletion of the wild-type allele of the PTEN/MMAC1 gene detected in the DNAs from hamartomas of various organs analyzed in Cowden's disease patients and from normal tissues adjacent to the tumors indicates that loss of its expression is an early event leading to increased proliferation, tissue disorganization, and susceptibility to tumor development (11,14). This contention is strongly supported by the recent observation that mice with germline heterozygous mutations in PTEN/MMAC1 develop increased susceptibility to tumors of various tissues (33-35).


    II. INACTIVATION OF PTEN/MMAC1 GENE IN MULTIPLE HUMAN TUMORS
 Top
 Abstract
 Introduction
 I. Germline PTEN/MMAC1 Mutations...
 II. Inactivation of PTEN/MMAC1...
 III. PTEN/MMAC1 as Tumor...
 Notes
 References
 
Besides the germline mutations in inherited hamartomatous syndromes, somatic mutations in the PTEN/MMAC1 gene are frequently found in a variety of sporadic human tumors in which the wild-type allele of the gene is inactivated mostly by deletions, thus conforming to the classical paradigm of tumor suppressor genes. The region on chromosome 10q that harbors the PTEN/MMAC1 gene and very likely other tumor suppressor gene(s) is hemizygously deleted in many human cancers, with a frequency reaching 60%-80% in the case of prostate cancer, endometrial carcinoma, and advanced glial tumors. Studies carried out within the last year have established PTEN/MMAC1 as the likely target gene of these deletions in many human tumors. The gene is inactivated by multiple mechanisms, including homozygous genomic deletions. In tumors with hemizygous deletions at chromosome 10q23, frameshift or nonsense mutations or mutations resulting in splice variants prematurely terminate the open reading frame of the remaining copy of the gene, thus producing a truncated and nonfunctional protein. Alternatively, the remaining copy of the gene is altered by missense mutations that are predicted to severely impair the function of its protein product. A detailed analysis of the nature and frequency of PTEN/MMAC1 gene mutations in different cancers may provide insights into the underlying genetic mechanisms of tumorigenesis in specific tissue types.

Endometrial Cancer

Endometrial carcinoma is the second most common noncolonic malignancy in hereditary nonpolyposis colon cancer (HNPCC) families, whose members carry germline mutations in mismatch repair genes resulting in replication errors and microsatellite instability (36). Among sporadic endometrial cancers, microsatellite instability is present only at a frequency of 20% (37). The two histologic variants of endometrial carcinoma, the endometrioid and serous types, appear to carry distinct genetic lesions. The less common but more aggressive serous carcinoma harbors mutations in the p53 gene with about 90% frequency, whereas K-ras mutations are restricted to the endometrioid type and are present in 20% of the tumors (38,39).

Several reports (39-45) show that PTEN/MMAC1 is the most frequently mutated gene identified yet in endometrial cancers. Inactivation of the PTEN/MMAC1 gene appears to be confined to tumors of endometrioid histology. In six different studies analyzing a total of 286 tumors (39-45), PTEN/MMAC1 mutations were detected in 33%-55% of the endometrial cancers, the highest frequency among the different tumor types examined so far (Table 1)Go. Furthermore, PTEN/MMAC1 mutations occur in early well-differentiated lesions (stage I/grade 1) as well as in very advanced and invasive tumors (stage IV/grade 3), suggesting their involvement in the initiation of endometrial tumorigenesis. That the inactivation of the PTEN/MMAC1 gene is an early event in endometrial carcinogenesis is also supported by data from two independent studies (46,47) showing a 22% frequency of PTEN/MMAC1 mutations in premalignant lesions of endometrial hyperplasia both with and without atypia.


View this table:
[in this window]
[in a new window]
 
Table 1. Mutations of the PTEN/MMAC1 gene in tumors of various tissue types*

 
Other striking features of the mutational profile of PTEN/MMAC1 in endometrial cancer are the substantially high frequency of mutations in microsatellite instability-positive tumors (78%-86%) compared with the microsatellite instability-negative tumors ({approx}30%) as well as the relative abundance of frameshift mutations (Fig. 2,Go A). It is not clear if the elevated frequency of frameshift mutations in endometrial carcinoma is a manifestation of the tumor's microsatellite instability-positive phenotype. A causal relationship is believed to exist between defective mismatch repair systems and alterations in microsatellite sequences in HNPCC (36). DNA polymerase slippage events at microsatellite repeat sequences lead to misalignment of DNA strands, which, if not repaired, give rise to deletions or insertions, resulting in frameshift mutations (48). However, the mutations in the PTEN/MMAC1 gene in endometrial carcinoma do not appear to be a consequence of a generalized replication error phenotype, as suggested by the following observations: 1) The frequency of mutations in the known DNA repair genes in endometrial carcinomas appears to be very low, 2) the repeat sequence in the transforming growth factor receptor BII gene, which is almost always mutated in microsatellite instability-positive colon cancers, is rarely mutated in endometrial cancers (49), and 3) the mutational spectrum of the PTEN/MMAC1 gene in microsatellite instability-negative endometrioid tumors does not differ from that in microsatellite instability-positive tumors (40,41). These observations argue that the PTEN/MMAC1 gene may be specifically targeted for mutations in endometrial carcinoma and is necessary for tumorigenesis.



View larger version (26K):
[in this window]
[in a new window]
 
Fig. 2. Mutational spectra of PTEN/MMAC1 in human cancers. Mutations reported in various tumors are compiled (see Table 1Go for references). Red = frameshift mutations; green = missense mutations; and blue = nonsense mutations. a) Mutation spectrum in endometrial carcinoma. b) Mutation spectrum in brain tumors. c) Mutations in various tumors are as indicated: B = breast; G = gastric; HN = head and neck squamous cell carcinoma; K = kidney; Lu = lung; Ly = lymphoma; M = melanoma; O = ovarian cancer; P = prostate; and T = thyroid. d) Spectrum of missense mutations. e) Spectrum of nonsense mutations. f) Spectrum of frameshift mutations. Mutations in d-f are pooled from a-c to present an overall representation of missense, nonsense, and frameshift mutations in the PTEN/MMAC1 gene. Black and red arrows under the bar = splice-site variants generated by point mutations and frameshift mutations, respectively. Structural and functional domains of the PTEN/MMAC1 gene are described in the legend for Fig. 1Go.

 
Malignant Glial Tumors

Evolution of glioblastomas, the highest malignancy grade glial tumors of the brain, is believed to occur via multiple genetic pathways. Notably, two pathways for which the molecular details are well characterized (50) are readily discernible in the formation of malignant gliomas. One subset of glioblastomas evolves from low-grade gliomas, which commonly harbor p53 mutations, via a progression pathway accumulating multiple genetic lesions that often include loss of chromosome 10 sequences. Alternatively, glioblastomas can develop in a de novo fashion following epidermal growth factor receptor amplification and gene losses on chromosome 10. Thus, loss of heterozygosity on chromosome 10 is a characteristic feature of high-malignancy-grade gliomas arising via both genetic pathways and occurs in 60%-80% of the tumors (50). PTEN/MMAC1 appears to be one of the target genes on chromosome 10 in glioblastomas and is found to be mutated in combination with either p53 mutations or epidermal growth factor receptor amplification (51). However, in many glioblastomas, the PTEN/MMAC1 gene is inactivated without accompanying alterations in either p53 or epidermal growth factor receptor genes. Thus, amplification of epidermal growth factor receptor, which harbors a receptor tyrosine kinase activity, and mutations in PTEN/MMAC1, with an associated phosphatase activity, appear to be independent of each other (51,52).

A total of 674 malignant glial tumors (anaplastic astrocytomas and glioblastomas multiforme) were analyzed, and PTEN/MMAC1 mutations were reported in 159 of these tumors, i.e., at a frequency of 24% (1,2,51-63). That PTEN/MMAC1 is not the only target for inactivation on chromosome 10 in gliomas is suggested by the fact that a considerable fraction of tumors with allelic losses on chromosome 10 retain a normal copy of the gene. Deletion mapping as well as functional studies have provided evidence for the presence of multiple tumor suppressor genes on both arms of chromosome 10 that may be involved in the development of glial neoplasms of different malignancy grades (64). Another putative tumor suppressor gene, DMBT1 (deleted in malignant brain tumors) with homology to the scavenger receptor cysteine-rich superfamily, has been identified on chromosome 10q25.3-26.1, which appears to be homozygously deleted in a subset of glioblastomas (65).

The spectrum of somatic mutations in the PTEN/MMAC1 gene in glioblastomas encompasses insertions, deletions, and point mutations, with the latter constituting about 60% of total mutations (Fig. 2,Go B). It is not clear if this mutational profile is attributable to alterations in the mismatch repair genes. Glioblastomas have been reported to be weakly microsatellite instability positive (66), and an association appears to exist between reduced levels of expression of multiple mismatch repair genes and tumor progression (67). It has been reported that MSH6, along with MSH2, functions in repairing single-base mismatches and smaller insertions/deletions and that cells derived from MSH6-deficient mice have a single nucleotide mismatch repair defect (68). Molecular profiling of the mismatch repair genes, in particular MSH6, in glioblastomas would be helpful in understanding the mutational spectrum of PTEN/MMAC1 with a relative abundance of base substitutions, a feature characteristic so far of the p53 gene, as opposed to frameshift mutations.

Prostate Carcinoma

Deletion mapping analysis of chromosome 10 in prostate carcinoma identified a complex pattern of allelic losses that includes 10q23.3, the region where the PTEN/MMAC1 gene resides (69). Analysis of 200 primary prostate carcinomas and 71 metastases from multiple sites, including those from pelvic lymph nodes, revealed 16 homozygous deletions and 10 mutations (70-74) (Table 1Go; Fig. 2,Go C). Of these, 13 (18%) tumors with homozygous deletions or point mutations derived from 71 metastases. The rather low frequency of mutations (10%) in primary and metastatic prostate carcinoma, which included both homozygous deletions and intragenic mutations, could be due to several factors. First, loss of the PTEN/MMAC1 gene may be particularly relevant to the progression and metastasis of advanced prostate carcinomas, a conclusion suggested by two studies (70,71). Second, tumors were not always screened for homozygous deletions and there may be a higher frequency of homozygous deletions than reported. Finally, the possibility exists that inactivation of the PTEN/MMAC1 gene occurs by mechanisms other than deletions and mutations. In this context, loss of expression or reduced expression of PTEN/MMAC1, both at the messenger RNA and protein levels, was recently demonstrated in xenografted prostate tumors (75). Restoration of PTEN/MMAC1 expression by demethylation in cells derived from the xenografts suggests loss of expression by promoter methylation (75). However, this mode of inactivation requires further exploration, since promoter methylation of the PTEN/MMAC1 gene could not be detected in prostate tumors with allelic loss of the gene and wild-type sequence of the remaining copy (70).

Ovarian Tumors

Epithelial ovarian tumors exist as four major histologic types (endometrioid, serous, mucinous, and clear cell), which probably evolve via distinct molecular pathways. Analysis of all four histologic types of ovarian tumors for the loss of heterozygosity on chromosome 10 and mutations in the PTEN/MMAC1 gene indicated that the gene is mutated predominantly, if not exclusively, in ovarian tumors of endometrioid origin (76,77). Although allelic losses on chromosome 10 were common in both endometrioid (43%) and serous (28%) tumors, PTEN/MMAC1 mutations were detected only in endometrioid ovarian tumors with a 26% frequency (Table 1Go) and in one of the 10 mucinous tumors examined that displayed a mixed mucinous and endometrioid histology. Thus, an important observation is that PTEN/MMAC1 mutations are common in endometrial as well as ovarian tumors of endometrioid histology and that tumors of both types containing such mutations are well or moderately differentiated, suggesting the involvement of PTEN/MMAC1 tumor suppressor function in disease initiation.

Other Cancers

Approximately 25%-50% of women with Cowden's disease develop breast cancers (5,6). It was, therefore, of interest to determine if PTEN/MMAC1 is involved in the pathogenesis of sporadic breast tumors. Screening of the PTEN/MMAC1 gene in 116 primary breast tumors, most of which included stage I and II tumors, showed an extremely low frequency of mutations (5%) (53,78,79) (Table 1Go). It is conceivable that PTEN/MMAC1 is inactivated mainly in advanced and metastatic breast tumors (79), as is the case in glial and prostatic tumors. It is, however, also possible that different subsets of genes are involved in sporadic versus inherited breast cancers, as is suggested by the absence of mutations in sporadic tumors of BRCA1 and BRCA2 genes responsible for inherited breast cancer in a number of pedigrees (80).

Thyroid cancer is also among the clinical manifestations of Cowden's disease. In an initial screen of 95 sporadic thyroid tumors encompassing many histologic types, one frameshift mutation in a papillary thyroid carcinoma was detected (81). It is possible that inactivation of the gene is associated with a particular histologic type. Alternatively, the gene may be inactivated in some tumor types by mechanisms other than mutations in the coding region.

Genetic alterations in the PTEN/MMAC1 gene were reported in seven of 67 head and neck squamous cell carcinomas (53,82,83), in one of eight kidney cancers (53), in four of 99 lung carcinomas (53,84-86), in three of 27 melanomas (53,87), in one of 29 gastric carcinomas (41), and in three of 68 lymphomas (88,89) examined (Table 1Go). Although the PTEN/MMAC1 gene is very frequently inactivated in high-malignancy-grade gliomas, mutations in the gene were detected at a much lower frequency in many other types of brain tumors (52,55,56,59-61,63,90). Of the 427 brain tumors of various types examined, including low-grade astrocytomas, pilocytic astrocytomas, medulloblastomas, oligodendrogliomas, oligoastrocytomas, and meningiomas, PTEN/MMAC1 alterations were detected only in seven tumors. Similarly, PTEN/MMAC1 appears to play a role in the development of endometrial and ovarian tumors of endometrioid histology, but mutations were not detected in other tumors of the female genital tract, such as the serous type of ovarian and endometrial tumors and squamous cervical carcinoma (40,76,77,91). Thus, the data obtained so far suggest that inactivation of the PTEN/MMAC1 gene is relevant in the formation of tumors of specific histology and tissue type.

Distribution of Mutations in PTEN/MMAC1 Gene

We have compiled PTEN/MMAC1 mutations in a variety of cancers reported through May 1999. In a total of about 2300 tumors analyzed and sequenced, 374 inactivating mutations, which include 44 homozygous deletions, have been reported (Table 1Go). Many tumor types were not examined specifically for homozygous deletions of the PTEN/MMAC1 gene or for mutations in the regulatory sequences; therefore, the overall mutation frequency of 16% may be an underestimation. Although the mutations are scattered over the entire gene, clustering of missense mutations in exons 5 and 6 (Fig. 2,Go D), of nonsense mutations in exon 7 (Fig. 2,Go E), and of frameshift mutations in and around the poly(A)6 stretches and phosphorylation sites in exon 7 and especially in exon 8 (Fig. 2,Go F) is obvious. The phosphatase domain located in exon 5 is the target of missense mutations. Most striking is the observation that the phosphatase core motif that constitutes less than 3% of the gene is the site of about 31% and 23% of point mutations reported so far in germline and sporadic tumors, respectively. A vast majority of germline mutations (74%) are point mutations compared with 51% of the somatic mutations detected in various tumors (two-sided P<.00003; Fisher's exact test ); in particular, nonsense mutations are substantially elevated in the germline. In contrast, 49% of the inactivating mutations in the PTEN/MMAC1 gene in tumors were frameshift mutations versus 26% in the germline (Table 2)Go.


View this table:
[in this window]
[in a new window]
 
Table 2. Summary of germline and somatic mutations in the PTEN/MMAC1 gene*

 
Among the missense mutations, transitions and transversions occur with about equal frequency. Altogether, 50 base substitutions in tumors result in a termination codon, and 22 of these (i.e., 44%) occur in three codons containing the dinucleotide CpG. There are nine codons containing CpG dinucleotides in the PTEN/MMAC1 open reading frame, three of which occur as CGA at codons 130, 233, and 335. The cytosine residue in these three codons is frequently mutated to thymine, thus terminating the open reading frame at the stop codon. Ten tumors had mutations at codon 130, which is located in the phosphatase core motif, and eight and four tumors each were mutated at codons 233 and 335, respectively. Another CpG dinucleotide present in codon 173 as CGC is mutated in 12 tumors; in eight of those tumors, the cytosine is replaced by thymine, resulting in the substitution of arginine for cysteine. This is consistent with the observation that frequent methylation of cytosine residues in CpG dinucleotides renders them prone to spontaneous deamination and a consequent C to T transition that occurs at a much higher rate than in unmethylated bases (92). The spectrum of point mutations in human tumors as well as in the germline also indicates that about one third of all base substitutions occur at a CpG dinucleotide.

Most of the PTEN/MMAC1 mutations were detected in glioblastomas and endometrial, ovarian, and prostate carcinomas. The mutational spectrum appears to have characteristic features for the tumors of each tissue type. In endometrial carcinoma, 60% of the mutations are frameshift versus 37% in glioblastomas (two-sided P<.0004; Fisher's exact test) (Table 2Go). Most of these frameshift mutations are 1-4 base pairs (bp) in length and occur predominantly in the N-terminal region of the gene or in exons 7 and 8, which contain potential tyrosine and/or serine phosphorylation sites (Fig. 2,Go A). Most notably, 15 single-base frameshift mutations (seven insertions and eight deletions of A) are clustered within the poly(A)6 stretch in codons 321-323 and nine tumors sustained a loss of 4 bp at the palindromic repeat sequence (TACTTACTTTA) at codons 317-320 (Fig. 2,Go A). Similarly, at another poly(A)6 stretch between codons 265-267, there was a loss of a single A in six tumors. One of the mechanisms for the generation of short deletions and insertions involves mispairing of the DNA strands at repeat sequences. Mononucleotide runs are believed to be the site of slipped misalignment of the template DNA strand, resulting in deletion or insertion if the nucleotide(s) excluded from pairing is on the template or primer strand, respectively. There is a clustering of missense mutations (23 [77%] of 30 missense mutations) in exon 5 and nonsense mutations (10 [45%] of 22 nonsense mutations) in exon 7 (Fig. 2,Go A).

The mutational spectrum of PTEN/MMAC1 in glioblastomas is distinct from that in endometrial carcinomas. In comparison with endometrial tumors, glioblastomas exhibit substantially fewer frameshift alterations but more missense mutations (Table 2Go; Fig. 2,Go B). This may reflect a different role for PTEN/MMAC1 in these two tumor types and/or differences in the mechanisms of mutation. Of the 150 mutations reported, 94 are single-base substitutions (i.e., 63% versus 40% in endometrial carcinoma), generating 61 missense, 21 nonsense, and 12 splice site variants (Table 2Go). Conspicuously, 17 mutations are clustered between codons 170 and codon 173 in exon 6, 10 of which (one frameshift and nine missense mutations) target codon 173, which contains one of the nine CpG sites of the gene and encodes for a highly conserved residue.

Mutations reported in various other cancers are shown in Fig. 2,Go C. Although the numbers are small, about 72% of the mutations (26 of 36) result in truncation of the protein. Five of the nine mutations reported in ovarian carcinoma are in the phosphatase core motif, four of them targeting codon 130.

The codons most frequently and differentially targeted by both germline as well as somatic mutations are listed in Table 3.Go Codons 129 and 130, which are located in the phosphatase core motif, are mutated at much higher frequencies in the germline than in various tumors. Codon 157, which is probably important for stabilizing the tertiary structure of the protein, is mutated in four Cowden's disease families but is not altered in any of the tumors. Of the four codons containing CpG dinucleotides that are frequently mutated in the germline and/or tumors, codon 130 is in the phosphatase core motif; 173 is conserved in tensin, auxilin, and bacterial phosphatase (2,3); and codons 233 and 335 are within or near tyrosine and/or serine phosphorylation sites. One striking feature is that codon 173 and its neighboring residues are predominantly mutated in glioblastomas with nine of the 12 point mutations occurring in glial tumors. Analysis of the enzymatic as well as tumor suppressor functions of the PTEN/MMAC1 protein mutated at codon 173 and elucidation of its signaling pathways in glial cells would be helpful in understanding the significance of mutational targeting of Arg 173 in glioblastomas.


View this table:
[in this window]
[in a new window]
 
Table 3. Frequently mutated codons in the PTEN/MMAC1 gene

 
The most obvious difference between the germline and somatic mutations is a very high density of somatic frameshift mutations at and around the two poly(A)6 stretches encompassing codons 265-267 and especially between codons 321 and 323, which is directly preceded by a palindromic sequence and a tyrosine phosphorylation site at codon 315. While there are differences between the germline and somatic mutations in terms of both the types and the frequency of mutations at specific codons/sites, the net effect in both cases is the generation of a truncated and probably nonfunctional or severely impaired protein.

Comparison of Mutational Spectra of Various Tumor Suppressor Genes With PTEN/MMAC1

Germline mutations in the PTEN/MMAC1 gene detected in the familial Cowden's disease and Bannayan-Zonana syndrome as well as somatic mutations in various tumors are distributed over the entire gene, with a clustering in three regions, i.e., phosphorylation sites in exons 7 and 8 and particularly in the phosphatase core motif in exon 5 (Figs. 1Go and 2Go). A great majority of these mutations, 62%-83%, are made up of frameshift, nonsense, and splice-site mutations, resulting in a truncated protein, a feature in common with the mutational spectra of other tumor suppressor genes. The germline mutations in BRCA1, BRCA2 (80), and patched (PTC) (93) genes are also scattered throughout these genes, and most are predicted to result in the truncation of the proteins. In particular, in the case of BRCA2, frameshift mutations constitute about 80% of the total mutations (94,95). A high density of the somatic adenomatous polyposis coli (APC) gene mutations occurs in the mutation cluster region, whereas the germline mutations are scattered throughout the 5' half of the gene (94). The mutational spectrum of the p53 gene is unique in the sense that a vast majority of p53 mutations are missense alterations (74%), with 22% of all mutations occurring in three hot spots on amino acids 175, 248, and 273 (94,95). A preponderance of missense mutations in the PTEN/MMAC1 gene (61 [41%] of a total of 150 mutations) is observed only in glioblastomas. The mutational spectrum of the PTEN/MMAC1 gene thus shares some common features with other tumor suppressor genes, such as distribution of mutations over the entire gene, clustering of mutations in three important functional domains, i.e., the phosphatase core motif in exon 5 and phosphorylation sites in exons 7 and 8, and a large number of missense mutations in some tumor types.


    III. PTEN/MMAC1 AS TUMOR SUPPRESSOR WITH DEFINED ENZYMATIC FUNCTION
 Top
 Abstract
 Introduction
 I. Germline PTEN/MMAC1 Mutations...
 II. Inactivation of PTEN/MMAC1...
 III. PTEN/MMAC1 as Tumor...
 Notes
 References
 
Structural and functional conservation is a hallmark of tumor suppressor genes. PTEN/MMAC1 exhibits remarkable structural similarity across species. The human gene is highly similar to yeast, mouse, and dog sequences, with a virtual sequence identity at the amino acid level between human and mouse genes (1-3). The protein product of the gene appears to have lipid phosphatase activity dephosphorylating the second messenger PIP3 (101-103). In the nematode Caenorhabditis elegans, DAF-18, a homologue of PTEN/MMAC1 functions in the insulin receptor-like signaling pathway. The role of DAF-18 is probably to decrease second messenger signaling by PIP3, thereby regulating metabolism, development, and lifespan of the worm (96,97). The conservation of signaling pathways between the nematodes and mammals raises the possibility that human PTEN/MMAC1, by virtue of its ability to modulate the phosphorylation of the second messenger-PIP3, may play a role not only in cancer development but also in aging and metabolic diseases (96,97).

Phosphorylation and dephosphorylation are key regulatory mechanisms in a variety of fundamental cellular processes, including cellular proliferation and differentiation. A critical balance among these processes is maintained by protein kinases, which are widely implicated in carcinogenesis, and protein phosphatases, which are the natural antagonists of the action of normal as well as oncogenic protein kinases and have been associated with cell cycle regulation and inhibition of cell proliferation (98,99). PTEN/MMAC1 is the first tumor suppressor gene identified in the phosphatase family (100), and the principal function of its gene product appears to be dephosphorylation of the second-messenger PIP3 (101-103). The growth-inhibitory function of PTEN/MMAC1 is evident when its wild-type complementary DNA is transfected into a variety of cancer cell lines, including glioblastoma, melanoma, and renal, breast, and prostate carcinomas (104-111). Mutant forms of PTEN/MMAC1 that retain protein phosphatase but lose lipid phosphatase activity fail to suppress the growth of glioma cell lines (107,110). The same mutations have been detected in Cowden's disease patients as well as in various sporadic tumors, providing strong evidence that lipid phosphatase activity is essential for its role in normal development and its tumor suppressor function. The growth-suppressive activity of PTEN/MMAC1 may be cell type specific and mediated by multiple mechanisms. In some cell lines, growth is suppressed by the ability of PTEN/MMAC1 to block cell cycle progression at the G1 phase by causing an increased synthesis of cyclin-dependent kinase inhibitor p27Kip1 (110,112,113), whereas, in other tumor cell lines, PTEN/MMAC1 has been shown to directly induce apoptosis (programmed cell death) (107,111,114-116).

The results in knockout mutant mice also demonstrate that the PTEN/MMAC1 plays an important role in both embryonic development and tumor suppression. Three independent knockout studies using different deletions in the PTEN/MMAC1 gene and mice with different genetic backgrounds showed differences in the details, such as timing of embryonic lethality, in the phenotype and in the spectrum of the lesions (34-36). One study (34) reported hyperplastic/dysplastic changes in the prostate, skin, and colon of PTEN/MMAC1 heterozygotes, resembling the features characteristic of Cowden's disease and Bannayan-Zonana syndrome. Neoplasms in multiple organs, including prostate, thyroid, gastrointestinal tract, and endometrium, were reported in another study (36), with multifocal endometrial complex atypical hyperplasia developing in all heterozygous females. This situation closely resembles human endometrial cancer, where PTEN/MMAC1 is the most commonly mutated gene and is involved in the early events of endometrial tumorigenesis with more than 20% of complex atypical hyperplasia containing mutations in the gene (46,47). Although leukemia is not a feature of Cowden's disease, a high incidence of T-cell lymphoma/leukemia was reported in two knockout studies (35,36).

Cells derived from PTEN/MMAC1-deficient mouse embryos contain elevated levels of intracellular PIP3 supporting a role of PTEN/MMAC1 in this signaling pathway (116). The intracellular levels of PIP3 directly control the phosphorylation of protein kinase B/Akt protooncogene, which is one of the key regulatory molecules of cell survival in response to multiple apoptotic stimuli (117). In PTEN/MMAC1-deficient mouse embryo fibroblasts, constitutively elevated activity of PKB/Akt together with decreased sensitivity to apoptosis suggests a role for PTEN/MMAC1 as a negative regulator of cell survival (116). An inverse association between the PTEN/MMAC1 protein and AKT/PKB was also detected in cell lines derived from hematologic malignancies (118).

Besides regulating the intracellular PIP3 levels, the phosphatase activity of the PTEN/MMAC1 protein might impact other cellular substrates/functions. In addition to modulating signal transduction and cell cycle progression, phosphorylation affects other cellular processes, such as cell-cell and cell-matrix adhesion, migration, and angiogenesis, processes that are important not only for normal cellular physiology but also impact tumor growth, invasion, and metastasis. The cytoplasmic location of the PTEN/MMAC1 protein (3) and its homology with tensin, which binds to actin filaments at focal adhesions and is implicated in the assembly of the signaling complexes, suggest that its physiologic substrates may reside in the cytoplasm. That PTEN/MMAC1 may regulate cellular signals originating from the cytoskeleton and/or focal adhesions is suggested by its inhibitory effects on integrin-mediated cell spreading, formation of actin-containing microfilaments and focal adhesions, and cell migration (119,120). These biologic activities of PTEN/MMAC1, at least in vitro, appear to be arbitrated by tyrosine dephosphorylation of the focal adhesion kinase, and the phosphatase domain of the gene is required for these functions (119,120). Inactivation of the PTEN/MMAC1 in human cancers may, therefore, contribute not only to their proliferation but also to their invasive potential.

Although the PTEN/MMAC1 is inactivated in multiple advanced cancers, it does show selectivity for specific target organs, as suggested by the absence or relative infrequency of mutations in various tumors. The pathway involving PTEN/MMAC1, PI3 kinase, and PKB/Akt appears to be critical in regulating cell survival in various tissues. A detailed characterization of this pathway and identification of other physiologic substrate(s)/mechanism(s) of action of PTEN/MMAC1 will be helpful in understanding the signaling pathways that are subverted not only during the progression of advanced cancers like glioblastomas and prostate carcinomas but also during the initiation of a large subset of endometrial and ovarian carcinomas.


    NOTES
 
We thank Dr. Barbara Dunn for her critical comments on the manuscript.


    REFERENCES
 Top
 Abstract
 Introduction
 I. Germline PTEN/MMAC1 Mutations...
 II. Inactivation of PTEN/MMAC1...
 III. PTEN/MMAC1 as Tumor...
 Notes
 References
 

1 Li J, Yen C, Liaw D, Podsypanina K, Bose S, Wang SI, Puc J, et al. PTEN, a putative protein tyrosine phosphatase gene mutated in human brain, breast, and prostate cancer. Science 1997;275:1943-7.[Abstract/Free Full Text]

2 Steck PA, Pershouse MA, Jasser SA, Yung WK, Lin H, Ligon A, et al. Identification of a candidate tumor suppressor gene, MMAC1, at chromosome 10q, 1997.23.3 that is mutated in multiple advanced cancers. Nature Genet 1997;15:356-62.[Medline]

3 Li DM, Sun Hl. TEP1, encoded by a candidate tumor suppressor locus, is a novel protein tyrosine phosphatase regulated by transforming growth factor b. Cancer Res 1997;57:2130-6.[Abstract]

4 Hanssen AM, Fryns JP. Cowden syndrome. J Med Genet 1995;32:117-9.[Abstract]

5 Eng C. Cowden syndrome. J Genet Counsel 1997;6:181-91.

6 Eng C, Murday V, Seal S, Mohammed S, Hodgson SV, Chaudary MA, et al. Cowden syndrome and Lhermitte-Duclos disease in a family: a single genetic syndrome with pleiotropy? J Med Genet 1994;31:458-61.[Abstract]

7 Ko FY, Wu TC, Hwang B. Intestinal polyps in children and adolescents: a review of 103 cases. Acta Paediatr 1995;36:197-202.

8 Fargnoli MC, Orlow SJ, Semelconcepcion J, Bolognia JL. Clinicopathologic findings in the Banayan-Riley-Ruvalcaba syndrome. Arch Dermatol 1996;132:1214-8.[Abstract]

9 Liaw D, Marsh DJ, Li J, Dahia PL, Wang SI, Zheng Z, et al. Germline mutations of the PTEN gene in Cowden disease, an inherited breast and thyroid cancer syndrome. Nat Genet 1997;16:64-7.[Medline]

10 Tsou HC, Teng DH, Ping XL, Brancolini V, Davis T, Hu R, et al. The role of MMAC1 mutations in early-onset breast cancer: causative in association with Cowden syndrome and excluded in BRCA1-negative cases. Am J Hum Genet 1997;61:1036-43.[Medline]

11 Lynch ED, Ostermeyer EA, Lee MK, Arena JF, Ji H, Dann J, et al. Inherited mutations in PTEN that are associated with breast cancer, cowden disease, and juvenile polyposis. Am J Hum Genet 1997;61:1254-60.[Medline]

12 Nelen MR, van Staveren WC, Peeters EA, Hassel MB, Gorlin RJ, Hamm H, et al. Germline mutations in the PTEN/MMAC1 gene in patients with Cowden disease. Hum Mol Genet 1997;6:1383-7.[Abstract/Free Full Text]

13 Marsh DJ, Dahia PL, Zheng Z, Liaw D, Parsons R, Gorlin RJ, et al. Germline mutations in PTEN are present in Bannayan-Zonana syndrome. Nat Genet 1997;16:333-4.[Medline]

14 Marsh DJ, Dahia PL, Coulon V, Zheng Z, Dorion-Bonnet, F, Call KM, et al. Allelic imbalance, including deletion of PTEN/MMAC1, at the Cowden disease locus on 10q22-23, in hamartomas from patients with Cowden syndrome and germline PTEN mutation. Genes Chromosomes Cancer 1998;21:61-9.[Medline]

15 Marsh DJ, Coulon V, Lunetta KL, Rocca-Serra P, Dahia PL, Zheng Z, et al. Mutation spectrum and genotype-phenotype analyses in Cowden disease and Bannayan-Zonana syndrome, two hamartoma syndromes with germline PTEN mutation. Hum Mol Genet 1998;7:507-15.[Abstract/Free Full Text]

16 Tsou HC, Ping XL, Xie XX, Gruener AC, Zhang HN, Nini R, et al. The genetic basis of Cowden's syndrome: three novel mutations in PTEN/MMAC1/TEP1. Hum Genet 1998;102:467-73.[Medline]

17 Scala S, Bruni P, Lo Muzio L, Mignogna M, Viglietto G, Fusco A. Novel mutation of the PTEN gene in an Italian Cowden's disease kindred. Int J Oncol 1998;13:665-8.[Medline]

18 Iida S, Nakamura Y, Fujii H, Kimura M, Moriwaki K. A heterozygous germline mutation of the PTEN/MMAC1 gene in a patient with Cowden disease. Int J Mol Med 1998;1:565-8.[Medline]

19 Raizis AM, Ferguson MM, Robinson BA, Atkinson CH, George PM. Identification of a novel PTEN mutation (L139X) in a patient with Cowden disease and Sjogren's syndrome. Mol Pathol 1998;51:339-41.[Abstract]

20 Stuphen R, Diamond TM, Minton SE, Peacocke M, Tsou HC, Root AW. Severe Lhermitte-Duclos disease with unique germline mutation of PTEN. Am J Med Genet 1999;82:290-3.[Medline]

21 Koch R, Scholz M, Nelen MR, Schwechheimer K, Epplen JT, Harders AG, et al. Lhermitte-Duclos disease as a component of Cowden's syndrome. Case report and review of literature. J Neurosurg 1999;90:776-9.[Medline]

22 Nelen MR, Kremer H, Konings IB, Schoute F, van Essen AJ, Koch R, et al. Novel PTEN mutations in patients with Cowden disease: absence of clear genotype-phenotype correlations. Eur J Hum Genet 1999;7:267-73.[Medline]

23 Zori RT, Marsh DJ, Graham GE, Marliss EB, Eng C. Germline PTEN mutation in a family with Cowden syndrome and Bannayan-Riley-Ruvalcaba syndrome. Am J Med Genet 1998; 80: 399-402.[Medline]

24 Celebi JT, Tsou HC, Chen FF, Zhang H, Ping XL, Lebwohl MG, et al. Phenotypic findings of Cowden syndrome and Bannayan-Zonana syndrome in a family associated with a single germline mutation in PTEN. J Med Genet 1999;36:360-4.[Abstract/Free Full Text]

25 Tok Celebi J, Chen FF, Zhang H, Ping XL, Tsou HC, Peacocke M. Identification of PTEN mutations in five families with Bannayan-Zonana syndrome. Exp Dermatol 1999;8:134-9.[Medline]

26 Olschwang, S, Serova-Sinilnikova OM, Lenoir GM, Thomas G. PTEN germ-line mutations in juvenile polyposis coli. Nat Genet 1998;18:12-4.[Medline]

27 Marsh DJ, Roth S, Lunetta KL, Hemminki A, Dahia PL, Sistonen P, et al. Exclusion of PTEN and 10q22-24 as the susceptibility locus for juvenile polyposis syndrome. Cancer Res 1997;57:5017-21.[Abstract]

28 Eng C, Peacocke M. PTEN and inherited hamartoma-cancer syndromes. Nat Genet 1998;19:223.[Medline]

29 Kurose K, Araki T, Matsunaka T, Takada Y, Emi M. Variant manifestation of Cowden disease in Japan: hamartomatous polyposis of the digestive tract with mutation of the PTEN gene. Am J Hum Genet 1999;64:308-10.[Medline]

30 Howe JR, Ringold JC, Summers RW, Mitros FA, Nishimura DY, Stone EM. A gene for familial juvenile polyposis maps to chromosome 18q21.1. Am J Hum Genet 1998;62:1129-36.[Medline]

31 Howe JR, Roth S, Ringold JC, Summers RW, Jarvinen HJ, Sistonen P, et al. Mutations in the SMAD4/DPC4 gene in juvenile polyposis. Science 1998;280:1086-8.[Abstract/Free Full Text]

32 Carethers JM, Furnari FB, Zigman AF, Lavine JE, Jones MC, Graham GE, et al. Absence of PTEN/MMAC1 germ-line mutations in sporadic Bannayan-Riley-Ruvalcaba syndrome. Cancer Res 1998;58:2724-6.[Abstract]

33 DiCristofano A, Pesce B, Cordon-Cardo C, Pandolfi PP. Pten is essential for embryonic development and tumour suppression. Nat Genet 1998;19:348-55.[Medline]

34 Suzuki A, de la Pompa JL, Stambolic V, Elia AJ, Sasaki T, del Barco Barrantes I, et al. High cancer susceptibility and embryonic lethality associated with mutation of the PTEN tumor suppressor gene in mice. Curr Biol 1998;8:1169-78.[Medline]

35 Podsypanina K, Ellenson LH, Nemes A, Gu J, Tamura M, Yamada KM, et al. Mutation of Pten/Mmac1 in mice causes neoplasia in multiple organ systems. Proc Natl Acad Sci U S A 1999; 96: 1563-8.[Abstract/Free Full Text]

36 Aaltonen LA, Peltomaki P, Mecklin JP, Jarvinen H, Jass JR, Green JS, et al. Replication errors in benign and malignant tumors from hereditary nonpolyposis colorectal cancer patients. Cancer Res 1994;54:1645-8.[Abstract]

37 Duggan BD, Felix JC, Muderspach LI, Tourgeman D, Zheng J, Shibata D. Microsatellite instability in sporadic endometrial carcinoma. J Natl Cancer Inst 1994;86:1216-21.[Abstract]

38 Enomoto T, Fujita M, Inoue M, Nomura T, Shroyer KR. Alteration of the p53 tumor suppressor gene and activation of c-K-ras-2 protooncogene in endometrial carcinoma from Colorado. Am J Clin Pathol 1995;103:224-30.[Medline]

39 Tashiro H, Isacson C, Levine R, Kurman RJ, Cho KR, Hedrick L. p53 gene mutations are common in uterine serous carcinoma and occur early in their pathogenesis. Am J Pathol 1997;150:177-85.[Abstract]

40 Tashiro H, Blazes MS, Wu R, Cho KR, Bose S, Wang SI, et al. Mutations in PTEN are frequent in endometrial carcinoma but rare in other common gynecological malignancies. Cancer Res 1997; 57: 3935-40.[Abstract]

41 Kong H, Suzuki A, Zou TT, Sakurada A, Kemp LW, Wakatsuki S, et al. PTEN1 is frequently mutated in primary endometrial carcinomas. Nat Genet 1997;17:143-4.[Medline]

42 Risinger JI, Hayes AK, Berchuck A, Barret JC. PTEN/MMAC1 mutations in endometrial cancers. Cancer Res 1997;57:4736-8.[Abstract]

43 Kurose K, Bando K, Fukino K, Sugisaki Y, Araki T, Emi M. Somatic mutations of the PTEN/MMAC1 gene in fifteen Japenese endometrial cancers: evidence for inactivation of both alleles. Jpn J Cancer Res 1998;89:842-8.[Medline]

44 Simpkins SB, Peiffer-Schneider S, Mutch DG, Gersell D, Goodfellow PJ. PTEN mutations in endometrial cancers with 10q LOH: additional evidence for the involvement of multiple tumor suppressors. Gynecol Oncol 1998;71:391-5.[Medline]

45 Risinger JI, Hayes K, Maxwell GL, Carney ME, Dodge RK, Barrett JC, et al. PTEN mutation in endometrial cancers is associated with favorable clinical and pathologic characteristics. Clin Cancer Res 1998;4:3005-10.[Abstract]

46 Maxwell GL, Risinger JI, Gumbs C, Shaw H, Bentley RC, Barrett C, et al. Mutation of the PTEN tumor suppressor gene in endometrial hyperplasias. Cancer Res 1998;58:2500-3.[Abstract]

47 Levine RL, Cargile CB, Blazes MS, van Rees B, Kurman RJ, Ellenson LH. PTEN mutations and microsatellite instability in complex atypical hyperplasia, a precursor lesion to uterine endometrioid carcinoma. Cancer Res 1998;58:3254-8.[Abstract]

48 Ripley LS. Frameshift mutation: determinants of specificity. Annu Rev Genet 1990;24:189-213.[Medline]

49 Myeroff LL, Parson R, Kim SJ, Hedrick L, Cho KR, Orth K, et al. A transforming growth factor beta receptor type II gene mutation common in colon and gastric but rare in endometrial cancers with microsatellite instability. Cancer Res 1995;55:5545-7.[Abstract]

50 Louis DN, Gusella JF. A tiger behind many doors: multiple genetic pathways to malignant glioma. Trends Genet 1995;11:412-5.[Medline]

51 Liu W, James CD, Frederick L, Alderete BE, Jenkins RB. PTEN/MMAC1 mutations and EGFR amplification in glioblastomas. Cancer Res 1997;57:5254-7.[Abstract]

52 Zhou XP, Li YJ, Hoang-Xuan K, Laurent-Puig P, Mokhtari K, Longy M, et al. Mutational analysis of the PTEN gene in gliomas: molecular and pathological correlations. Int J Cancer 1999;84:150-4.[Medline]

53 Teng DH, Hu R, Lin H, Davis T, Iliev D, Frye C, et al. MMAC1/PTEN mutations in primary tumor specimens and tumor cell lines. Cancer Res 1997;57:5221-5.[Abstract]

54 Wang SI, Puc J, Li J, Bruce JN, Cairns P, Sidransky D, et al. Somatic mutations of PTEN in glioblastoma multiforme. Cancer Res 1997;57:4183-6.[Abstract]

55 Rasheed BK, Stenzel TT, McLendon RE, Parsons R, Friedman AH, Friedman HS, et al. PTEN gene mutations are seen in high-grade but not in low-grade gliomas. Cancer Res 1997;57:4187-90.[Abstract]

56 Bostrom J, Cobbers JM, Wolter M, Tabatabai G, Weber RG, Lichter P, et al. Mutation of the PTEN (MMAC1) tumor suppressor gene in a subset of glioblastomas but not in meningiomas with loss of chromosome arm 10q. Cancer Res 1998;58:29-33.[Abstract]

57 Chiariello E, Roz L, Albarosa R, Magnani I, Finocchiaro G. PTEN/MMAC1 mutations in primary glioblastomas and short-term cultures of malignant gliomas.Oncogene 1998;16:541-5.[Medline]

58 Fults D, Pedone CA, Thompson GE, Uchiyama CM, Gumpper KL, Iliev D, et al. Microsatellite deletion mapping on chromosome 10q and mutation analysis of MMAC1, FAS, and MXI1 in human glioblastoma multiforme. Int J Oncol 1998;12:905-10.[Medline]

59 Duerr EM, Rollbrocker B, Hayashi Y, Peters N, Meyer-Puttlitz B, Louis DN, et al. PTEN mutations in gliomas and glioneuronal tumors. Oncogene1998 ;16:2259-64.[Medline]

60 Maier D, Zhang Z, Taylor E, Hamou MF, Gratzl O, Van Meir EG, et al. Somatic deletion mapping on chromosome 10 and sequence analysis of PTEN/MMAC1 point to the 10q25-26 region as the primary target in low-grade and high-grade gliomas. Oncogene 1998;16:3331-5.[Medline]

61 Watanabe K, Peraud A, Gratas C, Wakai S, Kleihues P, Ohgaki H. p53 and PTEN gene mutations in gemistocytic astrocytomas. Acta Neuropathol (Berl) 1998;95:559-64.[Medline]

62 Tohma Y, Gratas C, Biernat W, Peraud A, Fukuda M, Yonekawa Y, et al. PTEN (MMAC1) mutations are frequent in primary glioblastomas (de novo) but not in secondary glioblastomas. J Neuropathol Exp Neurol 1998;57:684-9.[Medline]

63 Davies MP, Gibbs FE, Halliwell N, Joyce KA, Roebuck MM, Rossi ML, et al. Mutation in the PTEN/MMAC1 gene in archival low grade and high grade gliomas. Br J Cancer 1999;79:1542-8.[Medline]

64 Ichimura K, Schmidt EE, Myakawa A, Goike HM, Collins VP. Distinct patterns of deletion on 10p and 10q suggest involvement of multiple tumor suppressor genes in the development of astrocytic gliomas of different malignancy grades. Genes Chromosomes Cancer 1998;22:9-15.[Medline]

65 Mollenhauer J, Wiemann S, Scheurlen W, Korn B, Hayashi Y, Wilgenbus KK, et al. DMBT1, a new member of the SRCR superfamily, on chromosome 10q25.3-26.1 is deleted in malignant brain tumours. Nat Genet 1997;17:32-9.[Medline]

66 Dams E, Van de Kleft EJ, Martin JJ, Verlooy J, Willems PJ. Instability of microsatellites in human gliomas. Cancer Res 1995;55:1547-9.[Abstract]

67 Wei Q, Bondy ML, Mao L, Guan Y, Cheng L, Cunningham J, et al. Reduced expression of mismatch repair genes measured by multiplex reverse transcription-polymerase chain reaction in human gliomas. Cancer Res 1997;57:1673-7.[Abstract]

68 Edelmann W, Yang K, Umar A, Heyer J, Lau K, Fan K, et al. Mutation in the mismatch repair gene Msh6 causes cancer susceptibility. Cell 1997;91:467-77.[Medline]

69 Ittmann M. Allelic loss on chromosome 10 in prostate adenocarcinoma. Cancer Res 1996;56:2143-7.[Abstract]

70 Cairns P, Okami K, Halachmi S, Halachmi N, Esteller M, Herman JG, et al. Frequent inactivation of TEN/MMAC1 in primary prostate cancer. Cancer Res 1997;57:4997-5000.[Abstract]

71 Suzuki H, Freije D, Nusskern DR, Okami K, Cairns P, Sidransky D, et al. Interfocal heterogeneity of PTEN/MMAC1 gene alterations in multiple metastatic prostate cancer tissues. Cancer Res 1998;58:204-9.[Abstract]

72 Feilotter HE, Nagai MA, Boag AH, Eng C, Mulligan LM. Analysis of PTEN and the 10q23 region in primary prostate carcinomas. Oncogene 1998;16:1743-8.[Medline]

73 Wang SI, Parsons R, Ittman M. Homozygous deletion of the PTEN tumor suppressor gene in a subset of prostate adenocarcinomas. Clin Cancer Res 1998;4:811-5.[Abstract]

74 Pesche S, Latil A, Muzeau F, Cussenot O, Fournier G, Longy M, et al. PTEN/MMAC1/TEP1 involvement in primary prostate cancers. Oncogene1998 ;16:2879-83.[Medline]

75 Whang YE, Wu X, Suzuki H, Reitter RE, Tran C, Vessella RL, et al. Inactivation of the tumor suppressor PTEN/MMAC1 in advanced human prostate cancer through loss of expression. Proc Natl Acad Sci U S A 1998;95:5246-50.[Abstract/Free Full Text]

76 Obata K, Morland SJ, Watson RH, Hitchcock A, Chenevix-Trench G, Thomas EJ, et al. Frequent PTEN/MMAC mutations in endometrioid but not serous or mucinous epithelial ovarian tumors. Cancer Res 1998;58:2095-7.[Abstract]

77 Yokomizo A, Tindall DJ, Hartmann L, Jenkins RB, Smith DI, Liu W. Mutation analysis of the putative tumor suppressor PTEN/MMAC1 in human ovarian cancer. Int J Oncol 1998;13:101-5.[Medline]

78 Rhei E, Kang L, Bogomolniy F, Federici MG, Borgen PI, Boyd J. Mutation analysis of the putative tumor suppressor gene PTEN/MMAC1 in primary breast carcinomas. Cancer Res 1997;57:3657-9.[Abstract]

79 Bose S, Wang SI, Terry MB, Hibshoosh H, Parsons R. Allelic loss of chromosome 10q23 is associated with tumor progression in breast carcinomas. Oncogene 1998;17:123-7.[Medline]

80 Casey G. The BRCA1 and BRCA2 breast cancer genes. Curr Opin Oncol 1997;9:88-93 .[Medline]

81 Dahia PL, Marsh DJ, Zheng Z, Zedenius J, Komminoth P, Frisk T, et al. Somatic deletions and mutations in the Cowden disease gene, PTEN, in sporadic thyroid tumors. Cancer Res 1997;57:4710-3.[Abstract]

82 Okami K, Wu L, Riggins G, Cairns P, Goggins M, Evron E, et al. Analysis of PTEN/MMAC1 alterations in aerodigestive tract tumors. Cancer Res1998 ;58:509-11.[Abstract]

83 Shao X, Tandon R, Samara G, Kanki H, Yano H, Close LG, et al. Mutational analysis of the PTEN gene in head and neck squamous cell carcinoma. Int J Cancer 1998;77:684-8.[Medline]

84 Kohno T, Takahashi M, Manda R, Yokota J. Inactivation of the PTEN/MMAC1/TEP1 gene in human lung cancers. Gene Chromosomes Cancer 1998;22:152-6.

85 Petersen S, Rudolf J, Bockmuhl U, Gellert K, Wolf G, Dietel M, et al. Distinct regions of allelic imbalance on chromosome 10q22-q26 in squamous cell carcinomas of the lung. Oncogene 1998;17:449-54.[Medline]

86 Yokomizo A, Tindall DJ, Drabkin H, Gemmill R, Franklin W, Yang P, et al. PTEN/MMAC1 mutations identified in small cell, but not in non-small cell lung cancers. Oncogene 1998;17:475-9.[Medline]

87 Tsao H, Zhang X, Benoit E, Haluska FG. Identification of PTEN/MMAC1 alterations in uncultured melanomas and melanoma cell lines. Oncogene 1998;16:3397-402.[Medline]

88 Nakahara Y, Nagai H, Kinoshita T, Uchida T, Hatano S, Murate R, et al. Mutational analysis of the PTEN/MMAC1 gene in non-Hodgkin's lymphoma. Leukemia 1998;12:1277-80.[Medline]

89 Gronbaek K, Zeuthen J, Guldberg P, Ralfkiaer E, Hou-Jensen K. Alterations of the MMAC1/PTEN gene in lymphoid malignancies. Blood 1998;91:4388-90.[Free Full Text]

90 Peters N, Wellenreuther R, Rollbrocker B, Hayashi Y, Meyer-Puttlitz B, Duerr EM, et al. Analysis of the PTEN gene in human meningiomas. Neuropathol App Neurobiol 1998;24:3-8.

91 Maxwell GL, Risinger JI, Tong B, Shaw H, Barrett JC, Berchuck A, et al. Mutation of the PTEN tumor suppressor gene is not a feature of ovarian cancers. Gynecol Oncol 1998;70:13-6.[Medline]

92 Bird AP. The relationship of DNA methylation to cancer. Cancer Surv 1996;28:87-101.[Medline]

93 Chidambaram A, Goldstein AM, Gailani MR, Gerrard B, Bale SJ, DiGiovanna JJ, et al. Mutations in the human homologue of the Drosophila patched gene in Caucasian and African-American nevoid basal cell carcinoma syndrome patients. Cancer Res 1996;56:4599-601.[Abstract]

94 Beroud C, Soussi T. p53 and APC gene mutations: software and databases. Nucleic Acids Res 1997;25:138.[Abstract/Free Full Text]

95 Hussain SP, Harris CC. Molecular epidemiology of human cancer: contribution of mutation spectra studies of tumor suppressor genes. Cancer Res 1998;58:4023-37.[Medline]

96 Ogg S, Ruvkun G. The C. elegans PTEN homolog, DAF-18, acts in the insulin receptor-like metabolic signaling pathway. Mol Cell 1998;2:887-93.[Medline]

97 Gil EB, Malone Link E, Liu LX, Johnson CD, Lees JA. Regulation of the insulin-like developmental pathway of Caenorhabditis elegans by a homolog of the PTEN tumor suppressor gene. Proc Natl Acad Sci U S A 1999;96:2925-30.[Abstract/Free Full Text]

98 Sun H, Tonks NK. The coordinated action of protein tyrosine phosphatases and kinases in cell signaling. Trends Biochem Sci 1994;19:480-5.[Medline]

99 Parsons R. Phosphatases and tumorigenesis. Curr Opin Oncol 1998;10:88-91.[Medline]

100 Meyers MP, Stolarov JP, Eng C, Li J, Wang SI, Wigler MH, et al. P-TEN, the tumor suppressor from human chromosome 10q23, is a dual-specificity phosphatase. Proc Natl Acad Sci U S A 1997;94: 9052-7.[Abstract/Free Full Text]

101 Maehama T, Dixon JE. The tumor suppressor, PTEN/MMAC1, dephosphorylates the lipid second messenger, phosphotidylinositol 3,4,5-triphosphate. J Biol Chem 1998;273:13375-8.[Abstract/Free Full Text]

102 Cantley LC, Neel BG. New insights into tumor suppression: PTEN suppresses tumor formation by restraining the phosphoinositide 3-kinase/AKT pathway. Proc Natl Acad Sci U S A 1999;96:4240-5.[Abstract/Free Full Text]

103 Maehama T, Dixon JE. PTEN: a tumour suppressor that functions as a phospholipid phosphatase. Trends Cell Biol 1999;9:125-8.[Medline]

104 Furnari FB, Lin H, Huang HS, Cavenee WK. Growth suppression of glioma cells by PTEN requires a functional phosphatase catalytic domain. Proc Natl Acad Sci U S A 1997;94:12479-84.[Abstract/Free Full Text]

105 Cheney IW, Johnson DE, Vaillancourt MT, Avanzini J, Morimoto A, Demers GW, et al. Suppression of tumorigenicity of glioblastoma cells by adenovirus-mediated MMAC1/PTEN gene transfer. Cancer Res 1998;58:2331-4.[Abstract]

106 Haas-Kogan D, Shalev N, Wong M, Mills G, Yount G, Stokoe D. Protein kinase B (PKB/AKT) activity is elevated in glioblastoma cells due to mutation of the tumor suppressor PTEN/MMAC. Curr Biol 1998;8:1195-8.[Medline]

107 Myers MP, Pass I, Batty IH, Van der Kaay J, Stolarov JP, Hemmings BA, et al. The lipid phosphatase activity of PTEN is critical for its tumor supressor function. Proc Natl Acad Sci U S A 1998;95:13513-8.[Abstract/Free Full Text]

108 Robertson GP, Furnari FB, Miele ME, Glendening MJ, Welch DR, Fountain JW, et al. In vitro loss of heterozygosity targets the PTEN/MMAC1 gene in melanoma. Proc Natl Acad Sci U S A 1998;95:9418-23.[Abstract/Free Full Text]

109 Davies MA, Lu Y, Sano T, Fang X, Tang P, LaPushin R, et al. Adenoviral transgene expression of MMAC/PTEN in human glioma cells inhibits Aktactivation and induces anoikis. Cancer Res 1998;58:5285-90.[Abstract]

110 Furnari FB, Huang HJ, Cavenee WK. The phosphoinositol phosphatase activity of PTEN mediates a serum-sensitive G1 growth arrest in glioma cells. Cancer Res 1998;58:5002-8.[Abstract]

111 Davies MA, Koul D, Dhesi H, Berman R, McDonnell TJ, McConkey D, et al. Regulation of Akt/PKB activity, cellular growth, and apoptosis in prostate carcinoma cells by MMAC/PTEN. Cancer Res 1999;59:2551-6.[Abstract/Free Full Text]

112 Li DM, Sun H. PTEN/MMAC1/TEP1 suppresses the tumorigenicity and induces G1 cell cycle arrest in human glioblastoma cells. Proc Natl Acad Sci U S A 1998;95:15406-11.[Abstract/Free Full Text]

113 Cheney IW, Neuteboom ST, Vaillancourt MT, Ramachandra M, Bookstein R. Adenovirus-mediated gene transfer of MMAC1/PTEN to glioblastoma cells inhibits S phase entry by the recruitment of p27 Kip1 into cyclin E/CDK2 complexes. Cancer Res 1999;59:2318-23.[Abstract/Free Full Text]

114 Li J, Simpson L, Takahashi M, Miliaresis C, Myers MP, Tonks N, et al. The PTEN/MMAC1 tumor suppressor induces cell death that is rescued by the AKT/protein kinase B oncogene. Cancer Res 1998;58:5667-72.[Abstract]

115 Wu X, Senechal K, Neshat MS, Whang YE, Sawyers CL. The PTEN/MMAC1 tumor suppressor phosphatase functions as a negative regulator of the phosphoinositide 3-kinase/Akt pathway. Proc Natl Acad Sci U S A 1998;95:15587-91.[Abstract/Free Full Text]

116 Stambolic V, Suzuki A, de la Pompa JL, Brothers GM, Mirtsos C, Sasaki T, et al. Negative regulation of PKB/Akt-dependent cell survival by the tumor suppressor PTEN. Cell 1998;95:29-39.[Medline]

117 Downward, J. Mechanisms and consequences of activation of protein kinase B/Akt. Curr Opin Cell Biol 1998;10:262-7.[Medline]

118 Dahia PL, Aguiar RC, Alberta J, Kum JB, Caron S, Sill H, et al. PTEN is inversely correlated with the cell survival factor Akt/PKB and is inactivated via multiple mechanisms in haematological malignancies. Hum Mol Genet 1999;8:185-93.[Abstract/Free Full Text]

119 Tamura M, Gu J, Matsumoto K, Aota S, Parsons R, Yamada KM. Inhibition of cell migration, spreading, and focal adhesions by tumor suppressor PTEN. Science 1998;280:1614-7.[Abstract/Free Full Text]

120 Tamura M, Gu J, Takino T, Yamada KM. Tumor suppressor PTEN inhibition of cell invasion, migration, and growth: differential involvement of focal adhesion kinase and p130cas. Cancer Res 1999;59:442-9.[Abstract/Free Full Text]

Manuscript received February 23, 1999; revised August 13, 1999; accepted September 21, 1999.


This article has been cited by other articles in HighWire Press-hosted journals:


             
Copyright © 1999 Oxford University Press (unless otherwise stated)
Oxford University Press Privacy Policy and Legal Statement