Tissue kallikreins are members of the S1 family (clan SA) of trypsin-like serine proteases and are present in at least six mammalian orders. In humans, tissue kallikreins (hK) are encoded by 15 structurally similar, steroid hormone–regulated genes (KLK) that colocalize to chromosome 19q13.4, representing the largest cluster of contiguous protease genes in the entire genome. hKs are widely expressed in diverse tissues and implicated in a range of normal physiologic functions from the regulation of blood pressure and electrolyte balance to tissue remodeling, prohormone processing, neural plasticity, and skin desquamation. Several lines of evidence suggest that hKs may be involved in cascade reactions and that cross-talk may exist with proteases of other catalytic classes. The proteolytic activity of hKs is regulated in several ways including zymogen activation, endogenous inhibitors, such as serpins, and via internal (auto)cleavage leading to inactivation. Dysregulated hK expression is associated with multiple diseases, primarily cancer. As a consequence, many kallikreins, in addition to hK3/PSA, have been identified as promising diagnostic and/or prognostic biomarkers for several cancer types, including ovarian, breast, and prostate. Recent data also suggest that hKs may be causally involved in carcinogenesis, particularly in tumor metastasis and invasion, and, thus, may represent attractive drug targets to consider for therapeutic intervention.

Proteases/peptidases, defined as enzymes that catalyze peptide bond hydrolysis, perform fundamental functions in all living organisms (1, 2). The “degradome” or complete set of proteases expressed at a given time within a cell, tissue, or organism (3) comprises ∼2% of all genes in many organisms. The human genome contains at least 553 protease genes and counting (4). Protease action is always irreversible and can involve either indiscriminant and non-specific degradation of protein substrates, as in apoptosis, or highly specific proteolytic processing or limited hydrolysis of selected target proteins, resulting in a functional change, as in prohormone activation. Proteases are classified according to three major criteria: location of the scissile peptide bond within the substrate (terminal or internal); catalytic mechanism; and evolutionary relationships, as revealed by structure. On the basis of the first criterion, proteases are broadly categorized as either exo- or endopeptidases, respectively. According to the second criterion, endopeptidases are divided into the well-known cysteine, serine, threonine, aspartic, and metalloprotease subgroups. Consistent with the third criterion, proteases of each catalytic class are clustered into several “clans,” which in turn include many “families” containing proteases that share significant sequence similarities (5, 6).

Serine proteases were among the first enzymes to be studied extensively (7). Their structural characteristics, catalytic mechanism, and roles in normal physiologic processes (e.g., digestion, coagulation, and cellular and humoral immunity) and in the pathology of many diseases (e.g., cancer, neurodegenerative disorders) have been previously reviewed (5, 8-12). With the exception of a small class of membrane-bound serine proteases, the vast majority are secreted. Furthermore, serine proteases have been organized into 11 evolutionary clans, most of which reside in clan SA of trypsin/chymotrypsin-like serine proteases (6).

Tissue kallikreins (EC 3.4.21) form a subgroup of secreted serine proteases within the S1 family of clan SA. To date, tissue kallikreins have been identified in a variety of species from six mammalian orders including

  1. Primates (e.g., human, chimpanzee, baboon, cynomolgus monkey, rhesus monkey, orangutan, gorilla);

  2. Rodentia (e.g., mouse, rat, guinea pig, mastomys);

  3. Carnivora (e.g., dog, cat);

  4. Proboscidea (e.g., elephant);

  5. Perissodactyla (e.g., horse); and

  6. Artiodactyla (e.g., pig, cow) (13-15).

The number of kallikreins varies among species from 2 in the dog (16) to more than 25 in rodents (15, 17). For a more thorough discussion of kallikreins in non-human species, please refer to our recent review (14). In humans, the tissue kallikrein (hK) family consists of 15 structurally homologous serine protease genes that colocalize in tandem to 19q13.4 (18). Of the ∼176 serine protease genes within the human genome, this family represents the largest contiguous cluster (4, 19). In fact, the kallikrein gene family is the largest cluster of protease genes of any catalytic class (4, 19).

The discovery of the complete hK family can be divided into two eras. The first era (1930s to 1980s) witnessed the discovery of the “classical” kallikreins. Although originally found in human urine (20, 21), human kallikrein 1 was subsequently identified at abundant levels in the pancreas (in Greek, the “kallikreas”), from which its name was derived (22). However, the gene for this kallikrein, now called KLK1, was not discovered until 1985 (23, 24). During the late 1980s, two genes with high structural similarity to KLK1, currently known as KLK2 and KLK3/PSA, were cloned and colocalized to the same chromosomal region (19q13.4) with KLK1 (25-27). At this time, it was concluded that the human kallikrein family had only three members, a statement that would hold for ∼10 years.

The second era (1994 to 2001) saw the expansion of the kallikrein family to 15 genes and the complete description of the human kallikrein locus. During the mid- to late 1990s, independent researchers cloned several novel serine protease genes with significant similarities to the classical kallikrein genes. These included, human stratum corneum chymotryptic enzyme (HSCCE)/KLK7 (28), normal epithelial cell-specific gene 1 (NES1)/KLK10 (29), protease M/zyme/neurosin/KLK6 (30-32), neuropsin/TADG-14/KLK8 (33, 34), and trypsin-like serine protease (TLSP)/hippostasin/KLK11 (35, 36). These genes were subsequently recognized as tissue kallikrein genes and mapped to 19q13.4 by Diamandis and colleagues (37-43). An additional seven kallikrein genes were independently cloned by our group and others, namely: prostase/KLK-L1/KLK4/ARM1/PRSS17 (44-46), human stratum corneum tryptic enzyme (HSCTE)/KLK-L2/KLK5 (47, 48), KLK-L3/KLK9 (40), KLK-L4/KLK13 (49), KLK-L5/KLK12 (50), KLK-L6/KLK14 (51, 52), and prostinogen/KLK15 (53, 54), as well as the first kallikrein pseudogene, ΨKLK1.2

2

Our data submitted for publication.

According to the official nomenclature, kallikrein gene and protein symbols are currently denoted “KLK” and “hK,” respectively (55). Gene numbering starts from centromere to telomere on chromosome 19q13.4 with the exception of the three classical kallikreins, for which the existing nomenclature was retained (56). Table 1 lists all official and alternative kallikrein gene and protein names. [Please note that plasma kallikrein (18, 57, 58) is not a member of the tissue kallikrein family.]

Table 1.

Official and Alternative Kallikrein Gene and Protein Names

Official Gene/ProteinOther Names/SymbolsGenBank AccessionsReferences
KLK1/hK1 Tissue/pancreatic/renal/urinary kallikrein, hPRK M25629, M33105 (24, 359) 
KLK2/hK2 Human glandular kallikrein 1, hGK-1 M18157 (25) 
KLK3/hK3 Prostate-specific antigen, PSA, APS X14810, M24543, M27274 (26, 360-362) 
KLK4/hK4 Prostase, KLK-L1, EMSP1, PRSS17, ARM1 AF113141 (44-46, 100, 229) 
KLK5/hK5 KLK-L2, HSCTE AF135028 (47-48) 
KLK6/hK6 Zyme, Protease M, Neurosin, PRSS9 D78203 (mRNA), AF149289 (Full gene) (30-32, 41) 
KLK7/hK7 HSCCE, PRSS6 L33404 (mRNA), AF166330 (Full gene) (28, 42) 
KLK8/hK8 Neuropsin, Ovasin, TADG-14, PRSS19, HNP AB009849 (33, 34) 
KLK9/hK9 KLK-L3 AF135026 (40) 
KLK10/hK10 NES1, PRSSL1 NM_002776 (mRNA), AF055481 (Full gene) (29, 39) 
KLK11/hK11 TLSP/Hippostasin, PRSS20 AB012917 (mRNA), AF164623 (Full gene) (35, 43) 
KLK12/hK12 KLK-L5 AF135025 (50) 
KLK13/hK13 KLK-L4 AF135024 (49) 
KLK14/hK14 KLK-L6 AF161221 (52) 
KLK15/hK15 Prostinogen, HSRNASPH AF242195 (53, 54) 
Official Gene/ProteinOther Names/SymbolsGenBank AccessionsReferences
KLK1/hK1 Tissue/pancreatic/renal/urinary kallikrein, hPRK M25629, M33105 (24, 359) 
KLK2/hK2 Human glandular kallikrein 1, hGK-1 M18157 (25) 
KLK3/hK3 Prostate-specific antigen, PSA, APS X14810, M24543, M27274 (26, 360-362) 
KLK4/hK4 Prostase, KLK-L1, EMSP1, PRSS17, ARM1 AF113141 (44-46, 100, 229) 
KLK5/hK5 KLK-L2, HSCTE AF135028 (47-48) 
KLK6/hK6 Zyme, Protease M, Neurosin, PRSS9 D78203 (mRNA), AF149289 (Full gene) (30-32, 41) 
KLK7/hK7 HSCCE, PRSS6 L33404 (mRNA), AF166330 (Full gene) (28, 42) 
KLK8/hK8 Neuropsin, Ovasin, TADG-14, PRSS19, HNP AB009849 (33, 34) 
KLK9/hK9 KLK-L3 AF135026 (40) 
KLK10/hK10 NES1, PRSSL1 NM_002776 (mRNA), AF055481 (Full gene) (29, 39) 
KLK11/hK11 TLSP/Hippostasin, PRSS20 AB012917 (mRNA), AF164623 (Full gene) (35, 43) 
KLK12/hK12 KLK-L5 AF135025 (50) 
KLK13/hK13 KLK-L4 AF135024 (49) 
KLK14/hK14 KLK-L6 AF161221 (52) 
KLK15/hK15 Prostinogen, HSRNASPH AF242195 (53, 54) 

NOTE: Adapted from ref. (18) with permission from copyright owners.

Thus far, the topology of the KLK locus has only been described in detail within the human (38, 59), chimpanzee,3

3

Our data submitted for publication.

mouse (17), and rat (15) genomes (Fig. 1). In general, mammalian KLK loci contain a single copy of KLK4-KLK15 genes and varying numbers of classical KLK genes and pseudogenes, all of which presumably arose due to gene duplication events. In humans, the KLK locus spans ∼300 kb on the long arm of chromosome 19 in cytogenic region 13.3 to 13.4 and is bound centromerically by the testicular acid phosphatase gene (ACPT; ref. 60), and telomerically by a cancer-associated gene (CAG)4
4

Our data submitted for publication.

and Siglec-9, a member of the sialic acid-binding Ig-like lectin (Siglec) family (61) (Fig. 1A). The KLK genes are tightly clustered in a tandem array, and are not intervened by any non-KLK genes. The three classical human KLK genes (KLK1, KLK2, and KLK3) are clustered within a 60-kb region with KLK15, whereas KLK4-KLK14 and the ΨKLK1 pseudogene are all located telomeric to KLK2. The direction of transcription of all genes is from telomere to centromere with the exception of KLK3 and KLK2.

FIGURE 1.

Organization of the tissue kallikrein gene loci in human (A), chimpanzee (B), mouse (C), and rat (D) genomes. Arrowheads indicate the location of genes and their direction of transcription. Green arrowheads, classical glandular kallikrein genes (KLK1, KLK2, and KLK3) and mouse and rat KLK1 paralogs. Blue arrowheads, non-classical kallikrein genes, KLK4KLK15. Kallikrein pseudogenes are represented by white arrowheads with the exception of rat KLK15 paralogous pseudogenes that are shown as striped white arrowheads. Grey arrowheads, non-kallikrein genes (ACPT, CAG, and Siglec-9). Official gene names are abbreviated to their numbers and indicated above each arrowhead. The nomenclature proposed by Olsson et al. (15) for paralogs of the mouse and rat loci is beneath each arrowhead. A. Gene lengths are indicated below each gene in the human kallikrein locus only. Figure is not drawn to scale. Modified from refs. (15, 17, 18) and our unpublished data (for chimpanzee).

FIGURE 1.

Organization of the tissue kallikrein gene loci in human (A), chimpanzee (B), mouse (C), and rat (D) genomes. Arrowheads indicate the location of genes and their direction of transcription. Green arrowheads, classical glandular kallikrein genes (KLK1, KLK2, and KLK3) and mouse and rat KLK1 paralogs. Blue arrowheads, non-classical kallikrein genes, KLK4KLK15. Kallikrein pseudogenes are represented by white arrowheads with the exception of rat KLK15 paralogous pseudogenes that are shown as striped white arrowheads. Grey arrowheads, non-kallikrein genes (ACPT, CAG, and Siglec-9). Official gene names are abbreviated to their numbers and indicated above each arrowhead. The nomenclature proposed by Olsson et al. (15) for paralogs of the mouse and rat loci is beneath each arrowhead. A. Gene lengths are indicated below each gene in the human kallikrein locus only. Figure is not drawn to scale. Modified from refs. (15, 17, 18) and our unpublished data (for chimpanzee).

Close modal

The syntenic relationship of KLK gene organization is relatively conserved among human, chimpanzee, and rodent genomes. The chimpanzee KLK locus is strikingly similar to the human locus, spans ∼350 kb of genomic sequence on chromosome 20, and contains orthologs to all 15 human KLK genes, which share more than 99% sequence similarity at the DNA and amino acid levels (Fig. 1B). Although the relative location and direction of transcription of KLK1 and KLK4-KLK14 is conserved among the human, chimpanzee, mouse, and rat loci, discrepancies exist with respect to locus size and number of KLK1 and KLK15 paralogs and pseudogenes within the mouse and rat genomes, due to additional gene duplication events. In the mouse, the KLK locus covers 590 kb within cytogenic region B2 on chromosome 7 and comprises 26 genes and 11 pseudogenes (Fig. 1C). In contrast to the human and chimp, the mouse lacks KLK2 and KLK3 orthologs but contains 23 KLK1 paralogs, 14 of which are functional. The latter reside between KLK1 and KLK15 in a 290-kb region, an area that is only 1.5 kb in length within the human locus. Generally, mouse and human KLK orthologs share ∼77% to 80% sequence similarity (62). The rat KLK locus spans 580 kb within cytogenic region q21 on chromosome 1, contains 22 genes and 19 pseudogenes, and is also devoid of KLK2 and KLK3 orthologs (Fig. 1D). Interestingly, this locus contains nine duplications of a ∼30-kb region harboring the KLK1, KLK15, and ΨKLK2 genes between ΨKLK2 and KLK4, resulting in nine paralogs of each gene. However, only the KLK1 paralogs are functional. Rat KLK genes share ∼80% to 85% sequence similarity with their human orthologs.

Phylogenetic analyses indicate that the classical KLKs of the human, mouse, and rat represent a distinct monophyletic group within the kallikrein family, separate from the more recently discovered kallikreins, KLK4-KLK15 (15, 17, 54, 59, 63). Moreover, within the classical KLK branch itself, KLK1 genes of the human, mouse, and rat form separate, species-specific subgroups, in contrast to rodent ΨKLK2 genes and human KLK2 and KLK3 that cluster into one subgroup. The latter suggests that these genes shared a common ancestral KLK2 gene that was subsequently silenced in rodents and evolved into KLK2 and KLK3 in primates (15). Taken together, the above evidence implies that the classical KLKs likely evolved independently in the human, mouse, and rat after the divergence of the lineages, in contrast to KLK4-KLK15 that probably diverged before their split. However, it is still not clear whether the classical KLKs have evolved from the KLK4-KLK15 group or if the two groups are monophyletic and share a common ancestor. In any event, the strong conservation of KLK4-KLK15 in these species suggests that the encoded serine proteases perform essential functions in mammals. In contrast, the “late-evolving” classical KLKs may encode proteins possessing functions that are rather unique to the primate and rodent orders (15).

Members of the human kallikrein gene family generally share 30% to 50% similarity at the DNA and amino acid levels, exclusive of KLK2 and KLK3 that possess 80% similarity and are, thus, considered the most closely related KLKs (18). As expected, KLK genes and the encoded hK proteins share several defining structural characteristics, illustrated and listed in Fig. 2. All KLK genes colocalize to the same chromosomal region (19q13.4) and code for serine proteases. KLK genes typically range from 4 to 10 kb in length and their organization, in terms of number and length of coding exons, intronic phase pattern, and location of start, stop and catalytic histidine, aspartate, and serine codons, is remarkably similar (18) (Fig. 2A). Of the few structural differences that have been found, all occur within non-coding regions. For instance, intron length is variable among KLK genes, leading to different gene lengths. Consensus GT-AG splice junctions are conserved among all KLKs, with the exception of the KLK10 gene that possesses a GC-AG splice site pair within intron 4 (39). The majority of the recently discovered KLKs (KLK4-KLK15) possess one or two non-coding exons within the 5′ untranslated region (UTR), whereas the classical KLKs (KLK1, KLK2, and KLK3) do not. Furthermore, the 3′ UTR beyond the stop codon is quite variable in length among KLKs and contains either a consensus (AATAAA) or variant polyadenylation signal ∼15 bp from the polyadenylic acid tail (18).

FIGURE 2.

Common structural features and schematic representation of a typical kallikrein gene (A) and protein (B). A.boxes, exons; lines, intervening intron. Coding exons are shown in red, green, and blue. Shaded boxes, untranslated exons and regions. The numbers above the exons indicate coding exon number and the Roman numerals below, the intron phase. Coding exon 1 harbors the start codon (indicated by ↓) and codes for the signal (red) and propeptides (green). Coding exons 2, 3, and 5 contain the histidine (H), aspartic acid (D), and serine (S) codons of the catalytic triad. Coding exon 5 harbors the stop codon (*). B.red box, signal peptide; green box, pro-peptide; blue box, mature, enzymatically active protein. Pro- and mature enzyme forms result from the sequential cleavage of the signal and pro-peptides on entry into the secretory pathway and on activation, respectively. It is important to note that the majority of these cleavage sites are predicted; only a few have been experimentally verified. Figure is not drawn to scale.

FIGURE 2.

Common structural features and schematic representation of a typical kallikrein gene (A) and protein (B). A.boxes, exons; lines, intervening intron. Coding exons are shown in red, green, and blue. Shaded boxes, untranslated exons and regions. The numbers above the exons indicate coding exon number and the Roman numerals below, the intron phase. Coding exon 1 harbors the start codon (indicated by ↓) and codes for the signal (red) and propeptides (green). Coding exons 2, 3, and 5 contain the histidine (H), aspartic acid (D), and serine (S) codons of the catalytic triad. Coding exon 5 harbors the stop codon (*). B.red box, signal peptide; green box, pro-peptide; blue box, mature, enzymatically active protein. Pro- and mature enzyme forms result from the sequential cleavage of the signal and pro-peptides on entry into the secretory pathway and on activation, respectively. It is important to note that the majority of these cleavage sites are predicted; only a few have been experimentally verified. Figure is not drawn to scale.

Close modal

Kallikrein proteins are single-chain secreted serine proteases translated as preproenzymes (Fig. 2B). Each contains a signal peptide of 16 to 30 amino acids at their NH2 terminus, followed by a pro-peptide of four to nine amino acids, and catalytic domain, which comprises the mature, enzymatically active, protein (18). Pro- and mature enzyme forms result from the sequential cleavage of the signal and pro-peptides on entry into the secretory pathway and on activation, respectively. It is important to note, however, that the majority of these cleavage sites are predicted; only a few have been experimentally verified (refs. 28, 36, 47, 64 and our unpublished data).

The calculated molecular weight of the peptide moiety of pro-hK proteins ranges from ∼23,000 to 26,000 (18). However, due to glycosylation, greater masses have been observed for several kallikreins including native hK1 (65), hK3 (66), hK5 (47), and hK7 (28) as well as recombinant hK2 (67) hK5 (47, 68), hK6 (69), hK7 (28, 70), and hK13 (71). (It is important to note, however, that the presence, absence, or type of glycosylation found in recombinant hKs produced in heterologous expression systems may not accurately reflect the status and nature of glycosylation of the native protein.) With the exception of hK1, in which O-linked glycosylation has been observed (65), all other reported glycosylation events, thus far, involve the addition of N-linked carbohydrates. Furthermore, glycosylation site prediction programs (72) on the Center for Biological Sequence Analysis website (http://www.cbs.dtu.dk) indicate that most hK proteins harbor one or more putative N-glycosylation sites or sequons, Asn-X-Ser/Thr (in which X is any amino acid except Pro; ref. 73), whereas only a few kallikreins have potential Ser/Thr residues involved in O-linked glycosylation. Collectively, experimental and bioinformatic data suggest that most, if not all, kallikreins are glycoproteins in vivo. Glycosylation of a number of proteins is important for their proper expression and function (74-77).

The amino acid of the S1 binding pocket, primarily responsible for substrate specificity in serine proteases (8), is found six amino acids NH2-terminal of the catalytic serine residue in all kallikrein enzymes at position 189, according to chymotrypsin numbering (78). Multiple alignments of deduced protein sequences indicate that 12 kallikreins possess an aspartate or glutamate residue in this position and are expected to cleave on the carboxyl side of basic amino acids such as arginine or lysine, similar to trypsin. In contrast, the remaining three, hK3, hK7, and hK9, have non-polar serine, asparagine, and glycine residues, respectively, conferring a chymotrypsin-like specificity (18) (Table 2). Thus far, experimental verification of substrate specificity has been done for all kallikrein enzymes with the exception of hK9, hK10, and hK12 (Table 2).

Table 2.

Specificity, Physiologic Substrates, and Post-translation Regulation of Human Kallikrein Proteins

hKPro-peptide Cleavage Site*S1 aaSpecificityP1 PositionPossible Physiologic SubstratesInhibitorsAuto ActivationActivation by Other hKAuto-/or Degradation
hK1 R↓I8 Asp Trypsin-like Arg, Met (363), Phe (364-366) Low MW kininogen (367), preANF (368), pro-insulin, LLP, prorenin, VIP, procollagenase, angiotensinogen (reviewed in ref. 191), BK B2 (369) Kallistatin (187, 188), PCI (370, 371), AAT (372, 373), placental bikunin (374)    
hK2 R↓I8 Asp Trypsin-like Arg (174, 365, 375) Seminogelin I/II (174, 204), IGFBP-3 (376), pro-uPA (243), fibronectin (204) PCI (174, 377), PI-6 (378), PAI-1 (379), ATIII (380), α2AP (381), ACT (185, 381), α2M (185, 377) ✓ (200, 201)  ✓ (173, 200) 
hK3 R↓I8 Ser Chymotrypsin-like Leu, Phe (382), Tyr (203) Seminogelin I/II, fibronectin (205, 233), laminin (233), lysozyme (382), plasminogen (248), IGFBP-3 (383, 384), TGF-β (294), PTHrp (385, 386) ACT (387, 388), α2M (387), PCI (389), AAT (390), ATIII (380)  hK2 (173, 202, 203), hK4 (242)§, hK15 (53) ✓ (391) 
hK4 Q↓I5 Asp Trypsin-like Arg (242),§, Lys (242) pro-uPA, PAP (242)  ✓ (242)   
hK5 R↓I38 Asp Trypsin-like Arg > Lys§ corneodesmosin (214), ECM, fibrinogen§ α2AP, ATIII, α2M§    
hK6 K↓L6 Asp Trypsin-like Arg > Lys (69, 392) ECM, fibrinogen, APP (178)∥, fibronectin, laminin (69), plasminogen ATIII, α2AP, AAT (178), ACT (178, 393) ✓ (31, 178)∥  ✓ (69, 178) 
hK7 K↓I8 Asn Chymotrypsin-like Tyr (28) IL-1β (218), corneodesmosin (214)    ✓ (28) 
hK8 K↓V5 Asp Trypsin-like Arg§ MBP (394)     
hK9 R↓A4 Gly Chymotrypsin-like       
hK10 R↓L10 Asp Trypsin-like       
hK11 R↓I4 Asp Trypsin-like Arg (36)§      
hK12 K↓I5 Asp Trypsin-like       
hK13 K↓V6 Asp Trypsin-like Arg > Lys§ ECM§, plasminogen (179) α2M, α2AP, ACT (71) ✓ (179)  ✓ (179) 
hK14 K↓I7 Asp Trypsin-like Arg > Lys§ ECM§    § 
hK15 K↓L6 Glu Trypsin-like Arg (53), Lys (53, 175)      
hKPro-peptide Cleavage Site*S1 aaSpecificityP1 PositionPossible Physiologic SubstratesInhibitorsAuto ActivationActivation by Other hKAuto-/or Degradation
hK1 R↓I8 Asp Trypsin-like Arg, Met (363), Phe (364-366) Low MW kininogen (367), preANF (368), pro-insulin, LLP, prorenin, VIP, procollagenase, angiotensinogen (reviewed in ref. 191), BK B2 (369) Kallistatin (187, 188), PCI (370, 371), AAT (372, 373), placental bikunin (374)    
hK2 R↓I8 Asp Trypsin-like Arg (174, 365, 375) Seminogelin I/II (174, 204), IGFBP-3 (376), pro-uPA (243), fibronectin (204) PCI (174, 377), PI-6 (378), PAI-1 (379), ATIII (380), α2AP (381), ACT (185, 381), α2M (185, 377) ✓ (200, 201)  ✓ (173, 200) 
hK3 R↓I8 Ser Chymotrypsin-like Leu, Phe (382), Tyr (203) Seminogelin I/II, fibronectin (205, 233), laminin (233), lysozyme (382), plasminogen (248), IGFBP-3 (383, 384), TGF-β (294), PTHrp (385, 386) ACT (387, 388), α2M (387), PCI (389), AAT (390), ATIII (380)  hK2 (173, 202, 203), hK4 (242)§, hK15 (53) ✓ (391) 
hK4 Q↓I5 Asp Trypsin-like Arg (242),§, Lys (242) pro-uPA, PAP (242)  ✓ (242)   
hK5 R↓I38 Asp Trypsin-like Arg > Lys§ corneodesmosin (214), ECM, fibrinogen§ α2AP, ATIII, α2M§    
hK6 K↓L6 Asp Trypsin-like Arg > Lys (69, 392) ECM, fibrinogen, APP (178)∥, fibronectin, laminin (69), plasminogen ATIII, α2AP, AAT (178), ACT (178, 393) ✓ (31, 178)∥  ✓ (69, 178) 
hK7 K↓I8 Asn Chymotrypsin-like Tyr (28) IL-1β (218), corneodesmosin (214)    ✓ (28) 
hK8 K↓V5 Asp Trypsin-like Arg§ MBP (394)     
hK9 R↓A4 Gly Chymotrypsin-like       
hK10 R↓L10 Asp Trypsin-like       
hK11 R↓I4 Asp Trypsin-like Arg (36)§      
hK12 K↓I5 Asp Trypsin-like       
hK13 K↓V6 Asp Trypsin-like Arg > Lys§ ECM§, plasminogen (179) α2M, α2AP, ACT (71) ✓ (179)  ✓ (179) 
hK14 K↓I7 Asp Trypsin-like Arg > Lys§ ECM§    § 
hK15 K↓L6 Glu Trypsin-like Arg (53), Lys (53, 175)      

Abbreviations: AAT, α1-antitrypsin; α2AP, α2-antiplasmin; APP, amyloid precursor protein; ATIII, antithrombin III; BK B2, human bradykinin B2 receptor; LLP, low density lipoprotein; MBP, myelin basic protein; PAI-1, plasminogen activator inhibitor-2; PAP, prostatic acid phosphatase; PCI, protein C inhibitor; PI-6, protease inhibitor-6; preANF, precursor of atrial natriuretic factor; pro-uPA, pro-form of urokinase-type plasminogen activator; PTHrp, parathyroid hormone-related peptide; VIP, vasoactive intestinal peptide.

*

Arrows indicate the cleavage site and pro-hK numbering is shown.

A chimeric form of hK4 (ch-hK4) was used.

§

Our unpublished data.

Our data submitted for publication.

G. Sotiropoulou, personal communication.

To date, X-ray crystallographic structures have been resolved for two human kallikreins, namely mature hK1 (79) and mature and pro-hK6 (64, 69), as well as for several non-human kallikreins, such as horse prostate kallikrein, an hK3 ortholog (80), mouse neuropsin/hK8 (81), mouse glandular kallikrein-13 (82), and porcine pancreatic kallikrein A (83). As members of the S1 family (clan SA) of serine proteases, kallikreins possess the archetypal tertiary structure of trypsin/chymotrypsin-like serine peptidases (6), which consist of two juxtaposed six-stranded anti-parallel β-barrels and two α-helices, with the active site [His57, Asp102, and Ser195, chymotrypsin numbering (78)] bridging the barrels (84, 85). The stereo ribbon plot for pro-hK6 is shown in Fig. 3.

FIGURE 3.

Crystal structure of pro-hK6 as solved by Gomis-Ruth et al. (64). Stereo ribbon plot of pro-hK6 shown in the traditional serine proteinase standard orientation (358) (i.e., looking into the active site cleft). The regular secondary structure elements are displayed as arrows (β-strands) and ribbons (α-helices) and labeled (β1-β12 and α1-α2). The side chains of the residues of the catalytic triad (light gray) and the six disulfide bonds (dark gray; SS1-SS6) are also shown as stick models and labeled. The NH2 and COOH termini and the positions of characteristic structural loops (i.e., autolysis loop, Ca-binding loop) are also indicated. (With dark gray coils are shown the poorly defined and undefined main-chain stretches.) Adapted with permission from ref. 64.

FIGURE 3.

Crystal structure of pro-hK6 as solved by Gomis-Ruth et al. (64). Stereo ribbon plot of pro-hK6 shown in the traditional serine proteinase standard orientation (358) (i.e., looking into the active site cleft). The regular secondary structure elements are displayed as arrows (β-strands) and ribbons (α-helices) and labeled (β1-β12 and α1-α2). The side chains of the residues of the catalytic triad (light gray) and the six disulfide bonds (dark gray; SS1-SS6) are also shown as stick models and labeled. The NH2 and COOH termini and the positions of characteristic structural loops (i.e., autolysis loop, Ca-binding loop) are also indicated. (With dark gray coils are shown the poorly defined and undefined main-chain stretches.) Adapted with permission from ref. 64.

Close modal

Structural heterogeneity among kallikrein enzymes can be attributed to the variable external surface loops surrounding the substrate-binding site, which are known to control their activity, define substrate and inhibitor specificity, and function in autolytic regulation (8, 69, 86-89). Depending on the hK in question, either five or six disulfide bonds serve to covalently link the polypeptide chain and provide structural rigidity to the surface loops surrounding the substrate-binding site. For instance, the classical kallikreins possess a unique surface loop named the “kallikrein loop,” not present in its entirety in any other kallikrein and absent in other serine proteases. Glycosylation of the kallikrein loop, and others, may serve to regulate kallikrein activity. For example, N-linked oligosaccharides present on the kallikrein loop determine the size of the S2 pocket and affect the P2 specificity of recombinant mouse hK8 (87). As well, the kallikrein loop, along with another surface loop, may be required for the regulated secretion of mouse hK8 (87). Within horse prostate kallikrein (hK3 ortholog), the kallikrein loop, seems to have a direct role in enzymatic control and substrate selectivity, because it protrudes over the catalytic region, blocking the entrance to the S1 specificity pocket (80). Furthermore, hK15 is unique in that it possesses an eight-amino-acid surface loop not present in any other kallikrein protein (54).

In the post-genomic era, with the discovery of an unexpectedly low number of genes (∼32,000) within the human genome sequence (90, 91), it has become clear that the generation of protein complexity occurs mainly via expansion of the human transcriptome. One of the major and most well-described mechanisms involved is alternative pre-mRNA splicing, whereby a single primary gene transcript or pre-mRNA gives rise to many mature mRNA transcripts, encoding many structurally and functionally distinct proteins (92). Indeed, recent genome-wide analyses indicate that 35% to 74% of all human genes have at least one alternative splice form (93, 94). The use of alternative promoters/transcriptional start sites (95) and polyadenylation signals (96) comprise additional sources for increasing the informational content of the genome.

Alternative pre-mRNA splicing, transcriptional start sites, and polyadenylation signals are common events among members of the kallikrein gene family. In addition to their classical mRNA forms, each kallikrein gene possesses at least one alternative transcript. In fact, a total of 70 alternative KLK mRNA isoforms have been identified to date, exclusive of the classical form (Table 3). Thus, a new dimension to the KLK family exists. With respect to alternative splicing events among KLK genes, the majority occur in coding regions and primarily involve exon skipping, followed by exon extension/truncation and intron retention, with only a few events occurring within the 5′ UTR. Consensus GT-AG splice sites are conserved in almost all kallikrein splice variants, with a few exceptions. For instance, a GC-AG splice site pair is present in a KLK5 variant with alternative splicing in the 5′ UTR (GenBank accession no. AY279381). A TG-AG splice junction is found in a KLK15 variant, in which coding exon 3 is lengthened and exon 4 is excluded (GenBank accession no. AY373374) and CC-AG pairs are found in several KLK3 variants (97). Recognition of these atypical splice sites by the spliceosome is possible in association with a conserved splice site (98, 99). Additional mRNA transcripts with alternative transcriptional start sites have been identified for several kallikrein genes including KLK3, KLK4, KLK5, KLK6, KLK7, and KLK11, and are likely the products of alternative promoters (refs. 26, 100-102 and our data submitted for publication). Furthermore, several of transcripts arising from alternative polyadenylation sites exist for KLK2 (103), KLK3 (97), and KLK7 (101). Many KLK transcripts exhibit a combination of alternative splicing events coupled with alternative transcription start site and polyadenylic acid signal usage.

Table 3.

Reported Human Kallikrein Variant Messenger RNA Transcripts

KallikreinReported No. of Variant Transcripts*GenBank Accession No.References
KLK1 NS (two variants), AY429508 refs. 395, 396 and our unpublished data 
KLK2 NS, AF188747, AF188746, AF188745, AY429510, AY429509 refs. 103, 139 and our unpublished data 
KLK3 11 AJ459783, AF335477, AF335478, AJ512346, AJ459784, M21896, AJ459782, AJ310937/M21897, AJ310938, NM_145864, NS (26, 97, 104, 105, 132, 397) 
KLK4 AF148532 (two variants), NS (two variants), AF228497, AF259964, AF259971, AF259970 (45, 100, 229, 398) 
KLK5 AY461805, AY279381, AY279380, AF435980, AF435981 ref. 101 and our unpublished data 
KLK6 AY279383, AY318867, AY318869, AY318870, AY318868, AY457039 our unpublished data 
KLK7 NM_19277, AF411215, AF411214 (42, 101) 
KLK8 NM_144505, NM_144506, NM_144506, BC040877 (102, 267, 399) 
KLK9 NS, AF135026 our unpublished data 
KLK10 BC002710 (399) 
KLK11 NM_144947, AB078780, BC022068/NM_006853 (36, 399, 400) 
KLK12 NM_019598, NM_145895, AY358524 (50, 401) 
KLK13 NS (five variants), AB108823, AB108824, AL050220 (108, 133) 
KLK14 NS (51) 
KLK15 AF242193 (three variants), AY373373, AY373374 ref. 54 and our unpublished data 
KallikreinReported No. of Variant Transcripts*GenBank Accession No.References
KLK1 NS (two variants), AY429508 refs. 395, 396 and our unpublished data 
KLK2 NS, AF188747, AF188746, AF188745, AY429510, AY429509 refs. 103, 139 and our unpublished data 
KLK3 11 AJ459783, AF335477, AF335478, AJ512346, AJ459784, M21896, AJ459782, AJ310937/M21897, AJ310938, NM_145864, NS (26, 97, 104, 105, 132, 397) 
KLK4 AF148532 (two variants), NS (two variants), AF228497, AF259964, AF259971, AF259970 (45, 100, 229, 398) 
KLK5 AY461805, AY279381, AY279380, AF435980, AF435981 ref. 101 and our unpublished data 
KLK6 AY279383, AY318867, AY318869, AY318870, AY318868, AY457039 our unpublished data 
KLK7 NM_19277, AF411215, AF411214 (42, 101) 
KLK8 NM_144505, NM_144506, NM_144506, BC040877 (102, 267, 399) 
KLK9 NS, AF135026 our unpublished data 
KLK10 BC002710 (399) 
KLK11 NM_144947, AB078780, BC022068/NM_006853 (36, 399, 400) 
KLK12 NM_019598, NM_145895, AY358524 (50, 401) 
KLK13 NS (five variants), AB108823, AB108824, AL050220 (108, 133) 
KLK14 NS (51) 
KLK15 AF242193 (three variants), AY373373, AY373374 ref. 54 and our unpublished data 
*

All mRNA transcripts (including splice variants, transcripts with alternative transcriptional start sites, and polyadenylation signals and combinations thereof) exclusive of the classical transcript.

Not submitted to GenBank.

By open reading frame analysis, it has been predicted that several alternatively spliced kallikrein transcripts will produce unique protein isoforms mainly due to in-frame usage of alternative translation initiation and termination codons, in-frame insertions or deletions in the middle of the protein sequence, and to a lesser extent, due to frameshifts that introduce premature stop codons. In most cases, the sequence encoding the signal peptide is retained, indicating that most kallikrein protein variants, on successful translation, are likely to be secreted and present in biological fluids, which may have clinical relevance in biomarker development. However, in the case of KLK4, one transcript isoform excludes the exon predicted to code for the signal peptide leading to the production of an intracellular protein (100), which may have unique functional implications. In some instances, alternative splicing may compromise the serine protease activity of the kallikrein protein due to exclusion of one or more residues of the conserved catalytic triad (H, D, S). Generally, most of these putative protein isoforms have not been isolated, with the exception of a few proteins encoded by KLK3 variants (97, 104, 105). Although the protein coding region is unaffected, variations in the 5′ or 3′ UTRs may have an effect on post-transcriptional regulation because these regions are known to be important in post-transcriptional regulation including mRNA stability, localization, and translational activation or repression (106, 107). Additional details on alternative kallikrein transcripts and their predicted proteins can be found in the literature cited in Table 3.

Kallikreins are expressed in a myriad of tissues at both the mRNA and protein levels. As delineated by Northern blot, reverse transcription-PCR, and ELISA methodologies collectively, each kallikrein displays a relatively broad tissue expression pattern, with highest expression levels within a few major tissues and lower levels of expression in many others (18, 63, 71, 108-116). Interestingly, kallikreins are often coexpressed within the same tissues. The most notable example is the concurrent and almost exclusive expression of KLK2, KLK3, KLK4, KLK11, and KLK15 in the prostate, at the mRNA level. As well, almost every kallikrein is expressed in the salivary gland, while subgroups reside in the skin (KLK1, KLK4, KLK5, KLK6, KLK7, KLK8, KLK9, KLK10, KLK11, KLK13, and KLK14), breast (KLK5, KLK6, KLK10, KLK13), pancreas (KLK1, KLK6-KLK13), and the central nervous system (KLK6, KLK7, KLK8, KLK9, KLK14). The presence of kallikreins in biological fluids, such as serum, seminal plasma, and the milk of lactating women, confirms that they are secreted proteins in vivo. The functional implications of kallikrein coexpression are discussed in a later section.

Furthermore, in situ and/or immunohistochemistry studies indicate that kallikreins, including hK3 (117, 118), hK4 (109, 119), hK6 (120-122), hK7 (123), hK9 (124), KLK10/hK10 (121, 122, 125), KLK11/hK11 (115, 126), hK13 (122, 127), KLK14/hK14 (51, 116), are localized predominantly in the cytoplasm of glandular epithelia, from which they are likely secreted. The hK1, hK6, hK10, and hK13 proteins have also been localized to the epithelium of the choroid plexus and other cell types within the central and peripheral nervous systems (120, 121, 127, 128). With respect to the skin, hK5 and hK7 expression was found to be restricted to the stratum granulosum of the normal epidermis (129-131). A recent in situ hybridization study indicates that several other KLKs are also prominently expressed in the stratum granulosum as well as in the inner root sheath of hair follicular epithelium and the cytoplasm of cells within the eccrine sweat glands and sebaceous glands (108).

Tissue-specific patterns of expression have also been documented for many alternative mRNA transcripts of kallikrein genes. For example, a KLK2 and KLK3 splice variant, both with a partially retained intron, are exclusively expressed in the prostatic epithelium (132). Splice variants of KLK4, KLK8, and KLK13 gene transcripts were found to be the predominant mRNA species in the skin (108). One variant of the KLK8 gene is predominately expressed in the pancreas, while another variant is preferentially expressed in adult brain and hippocampus (102). The KLK11 gene has two tissue-specific mRNA isoforms, known as the brain type and prostate type, the former of which is expressed in the brain and prostate and the latter that is expressed exclusively in the prostate (36). Furthermore, several testis-specific splice variants of KLK13 have been identified (133). Hooper et al. (51) have discovered a 1.5-kb transcript of KLK14 transcribed only in the prostate and another 1.9-kb transcript expressed exclusively in skeletal muscle. Furthermore, on transfection of a green fluorescent protein (GFP)-tagged KLK4 transcript variant, lacking the sequence coding for the signal peptide, into COS and HeLa cells, the encoded protein was predominantly localized in the nucleus (100). The physiologic and clinical relevance of alternative kallikrein mRNA transcripts warrants further investigation.

Transcriptional Control of Gene Expression

Transcriptional regulation of eukaryotic genes is a complex process that requires many basal transcription factors for initiation and promoter-specific regulatory protein(s) (activators or repressors) that either enhance or repress target gene expression depending on the nature of signaling stimuli (134).

The regulation of gene expression by steroid hormones, mediated on binding to their cognate receptors, plays an important role in the normal development and function of many organs as well as in the pathogenesis of endocrine-related cancers (135-138). Numerous in vitro and in vivo studies confirm that all human kallikrein genes are under steroid hormone regulation in endocrine-related tissues and cell lines (41, 42, 44, 48-50, 52, 54, 113, 139-147). The most notable example is the classical up-regulation of KLK2 and KLK3 transcription in response to androgens and progestins in prostate and breast cancer cell lines (139, 142, 143). Conversely, other kallikreins such as KLK1, KLK6, and KLK10 are more responsive to estrogens (41, 145, 148). An interesting observation is the differential pattern of hormonal regulation of certain genes, for instance, KLK4 is up-regulated by androgens in prostate and breast cancer cell lines (44, 46) and by estrogens in endometrial cancer cell lines (144) and KLK12 is up-regulated by androgens and progestins in prostate cancer cell lines and by estrogens and progestins in breast cancer cell lines (50).

Functional characterization of kallikrein gene promoters and enhancers may aid in delineating the mechanism of transcriptional regulation by steroid hormone-receptor complexes. These complexes can modulate transcription of target genes in a direct or indirect fashion (149). In the former, the complex binds directly to cis-acting DNA sequences known as hormone response elements (HRE) in the promoter/enhancer regions of regulated genes, thereby recruiting necessary cofactors that interact with the basic transcription machinery to regulate gene expression. In the indirect pathway, hormone-receptor complexes do not bind to cognate hormone response elements and indirectly modulate gene expression via interactions with trans-acting transcription factors. Thus far, promoters have only been characterized for KLK1-3 and KLK10.

The KLK1 promoter harbors a putative estrogen response element (ERE) thought to mediate estrogenic regulation, but has not been functionally tested (150). A number of androgen responsive elements (ARE) within the proximal promoter and enhancer regions of KLK2 and KLK3 genes have been identified and believed to be primarily responsible for transcriptional regulation by androgens. KLK2 has two AREs; one at position −170 within its promoter (25, 141) and another in the enhancer region −3819 to −3805 upstream from the transcription start site (151). The KLK3 proximal promoter harbors two functional AREs (ARE-I and ARE-II) at positions −170 and −400 (140, 152) and an additional ARE (ARE-III) in the far upstream enhancer region (−4,136), which has a dramatic effect on KLK3 transcription, in comparison to ARE-I and ARE-II (153-156). Furthermore, five additional low-affinity AREs have been identified close to ARE-III (157). Conversely, KLK10 promoter and enhancer regions do not harbor functional hormone response elements directly involved in mediating the apparent transcriptional regulation by steroid hormone-receptor complexes (146). As is the case for other genes and gene families, active hormone response elements may be located within exons or UTRs of the KLK10 gene or elsewhere in the kallikrein locus, respectively.

The promoter and enhancer regions of the 11 remaining human kallikrein genes have not as yet been functionally characterized. However, sequence analysis has identified putative AREs in the promoter regions of KLK4, KLK14, and KLK15 genes (45, 147, 158).

Accumulating reports indicate that the function of steroid hormone receptors is regulated by many coactivators/repressors that act as bridging molecules between hormone-receptor complexes and the basal transcription machinery to either activate or inhibit transcriptional regulation (159). For instance, the relative levels of several coactivators/repressors might differentially modulate the transcriptional activity within the promoter/enhancer region of KLK2 and KLK3 of various breast cancer cell lines (160).

Furthermore, several recent studies point to the possibility of cross-talk between steroid hormone signaling with other signal transduction pathways in the regulation of kallikrein gene transcription. For instance, Sadar (161) suggest that cross-talk between androgen receptor (AR) and protein kinase A signal transduction pathways contributes to the androgen-independent induction of KLK3 gene expression. Transcription factors activator protein and a Fos-containing protein complex distinct from activator protein were also reported to regulate KLK2 and KLK3 gene transcription (162, 163). As well, the KLK10 promoter was found to harbor potential AP1-binding, SP1-binding, and adenosine 3′,5′-monophosphate responsive element sequences (164). Wang et al. (165) have discovered that a novel GAGATA transcription factor binds to a cis-regulatory element located within the enhancer region of the KLK3 promoter and is required for the maximum transcriptional response to androgens. Conversely, a negative regulatory cis-element named XBE was identified within the KLK3 enhancer region, and was found to recruit both the AR and the p65 subunit of nuclear factor (NF)-κB AR, leading to the down-regulation of AR-mediated transcription of KLK3 (166). Thus, cross-talk exists between AR and NF-κB p65 transcription factors and was found to occur via novel mechanism through which the factors compete for binding at a common DNA element.

Moreover, epigenetic control of gene expression such as DNA methylation may also be implicated in regulation of kallikrein gene transcription, particularly during carcinogenesis. The dramatic-down regulation of the KLK10 gene in breast cancer and in acute lymphoblastic leukemia has been attributed primarily to hypermethylation of exon 3 with this gene (167, 168). This mechanism is also thought to explain, in part, KLK10 silencing in ovarian and prostate cancers.5

5

Our data submitted for publication.

Another mechanism of transcriptional control involves locus control regions, a class of cis-acting regulatory elements that regulate the expression of linked genes in a tissue and copy number-specific manner in a wide spectrum of mammalian gene families (169, 170), including rodent kallikrein gene families (171). Smith et al. (171) propose that a dominant locus control region controls the tissue-specific expression of all rat kallikrein genes in the salivary gland, in conjunction with gene-associated regulatory elements within promoter and enhancer regions. Given the above and the fact that all human KLK genes, except KLK2 and KLK3, are transcribed in the same direction (from telomere to centromere) and that many are coexpressed within tissues, locus control regions may also be implicated in the coordinate regulation and expression of human kallikrein genes.

Therefore, although steroid hormones play a major role, the control of kallikrein gene transcription may involve integration of a myriad of transcription factors and pathways, including epigenetic mechanisms and locus control regions, which serve to increase regulatory diversity and provide opportunities for cell and tissue-specific responses.

Post-translational Control of Protein Function

One of the main characteristics of proteases is their ability to catalyze reactions irreversibly. As a consequence, several mechanisms have evolved to spatially and temporally regulate serine protease activity to prevent unwanted protein degradation, including: (a) zymogen activation, (b) internal cleavage, and (c) endogenous inhibitors. First, all known proteases are synthesized as zymogens or inactive precursors, which possess an inhibitory pro-peptide that sterically blocks the active site and thereby prevents substrate binding. Zymogen conversion to the active enzyme generally occurs by limited proteolysis of the pro-peptide, via diverse mechanisms, including enzymatic or non-enzymatic cofactors that trigger activation, to a simple pH change resulting in autoactivation (reviewed in ref. 172) and can occur intracellularly (within the secretory pathway) or extracellularly. All pro-hKs, expect hK4, require the activity of a trypsin-like serine protease for activation as shown in Table 2. Thus, several studies have reported activation of pro-hKs by trypsin, enterokinase, trypsin-like hKs, and via autoactivation, in vitro (to be discussed in detail in the following section).

Once activated, serine proteases may be inactivated by internal cleavage followed by degradation. Cleavage may be autolytic or mediated by another protease. This mechanism has been reported for six members of the kallikrein family. For instance, degraded forms of hK2 with a major cleavage site between residues R145-S146 and a minor site between R101-L102 have been isolated from seminal plasma and prostate tissues (173, 174). As well, hK3 purified from the seminal plasma (∼30% of total hK3) and prostatic tissues contains internal peptide bond cleavages at one major, K145-K146, and two minor, R85-F86 and K182-S183 sites, leading to inactivation (175-177). The enzyme(s) responsible for internal cleavage of hK2 or hK3 are still unknown; however, hK2 is likely autodegraded because this enzyme possesses trypsin-like activity and the cleavage sites (P1-Arg) require a trypsin-like specificity. Furthermore, hK6 (178) and hK13 (179) are capable of autoinactivation in vitro, between residues R76-E77 and R114-S115, respectively. Self-digestion has also been reported for hK7 (28) and hK14,5

5

Our unpublished data.

yet the cleavage sites have not been determined.

Many endogenous inhibitors are known to regulate the activity of serine proteases. Laskowski et al. (180, 181). Generally speaking, many specific inhibitors are capable of inhibiting the same serine protease, and the same inhibitor may inhibit several serine proteases (180). Many kallikreins form complexes in vivo and/or in vitro with plasma inhibitors, primarily serpins and α2-macroglobulin (α2M) (Table 2). The interaction of serpins with serine proteases can occur via (a) the inhibitory pathway, leading to complex formation that results in the deformation and irreversible inactivation of the protease or (b) the substrate pathway, in which the serpin is cleaved by the protease and does not result in the inhibition of the protease (182, 183). hK1, 2, 3, 5, 6, and 13 form complexes and are inhibited in biological fluids such as serum, ascites fluid, seminal plasma, cerebrospinal fluid, milk of lactating women, by various serpins, including α1-antitrypsin (a.k.a. α1-protease inhibitor), α1-antichymotrypsin (ACT), protein C inhibitor (a.k.a. plasminogen activator inhibitor-3), plasminogen activator inhibitor-1, antithrombin III, and α2-antiplasmin. For instance, the major fraction (70% to 90%) of total hK3 in serum is complexed with ACT (184), whereas hK2 is complexed with ACT and protein C inhibitor in serum but at a much lower proportion (4% to 19%; refs. 185, 186) Interestingly, kallistatin is a uniquely specific serpin for hK1 (187, 188). Some kallikreins also interact with serpins via the substrate pathway, for example, hK3 with ACT (189) and hK5 and hK6 with α2AP (ref. 178 and our unpublished data). The interaction of proteases with α2M involves a molecular trap mechanism, which does not lead to inhibition of protease activity but prevents proteases from interacting with large substrates or inhibitors by steric hindrance (190). Human kallikreins 2, 3, 5, and 13 interact with α2M in the serum. Furthermore, it has been proposed that inhibitory peptides derived from serine protease inhibitor Kazal-type 5 (SPINK5) may regulate hK activity due to the colocalization of KLKs and SPINK5 in the skin (108).

Due to their presence in diverse tissues and cell types and by virtue of their serine protease activity, kallikreins are implicated in a wide range of normal physiologic processes, from the regulation of cell growth to tissue remodeling (Table 2). To date, several biological roles have been established for classical kallikreins, hK1, hK2, and hK3. The primary activity of hK1 involves the cleavage of low molecular weight kininogen to release lysyl-bradykinin (kallidin), which in turn binds to its receptors, bradykinin B1 and B2, in target tissues and mediates varied processes such as regulation of blood pressure, smooth muscle contraction, neutrophil chemotaxis and pain induction, vascular permeability, vascular cell growth, electrolyte balance, and inflammatory cascades (191, 192). A role for the hK1-kinin system in the establishment and maintenance of placental blood flow through vasodilation, platelet antiaggregation, cell proliferation, and trophoblast invasion during different stages of pregnancy has also been suggested (193, 194). Apart from its kininogenase activity, hK1 is implicated in other tissue and/or cell-specific functions, including processing growth factors and peptide hormones (listed in Table 2) in the pituitary, pancreas, and other tissues (191, 195-198).

Unlike hK1, both hK2 and hK3 possess relatively low kininogenase activity (199). Four in vitro studies have shown that hK2 is able to activate itself (67, 173, 200, 201). A series of contradictory studies has been published with respect to the activation of pro-hK3 by hK2 (173, 201-203). In 1997, three independent groups published that mature hK2 was able to activate pro-hK3, at a slow rate (173, 202, 203). However, in a subsequent experiment, Denmeade et al. (201) showed that hK2 was unable to cleave the fluorogenic pro-hK3 peptide substrate APLILSR-AMC, calling into question the previous findings. Once enzymatically active, hK2 and hK3 contribute to seminal clot liquefaction after ejaculation, which is integral to sperm motility, through their hydrolysis of seminal vesicle proteins, seminogelin I and II, and fibronectin (204, 205). However, hK3 cleaves these substrates at a higher efficiency and at different sites compared with hK2. As listed in Table 2, many other potential substrates, including growth factor binding proteins, peptide hormones, and components of the basement membrane/extracellular matrix (ECM), have been identified for hK2 and hK3.

Thus far, the physiologic roles of the remaining kallikreins, hK4 through hK15, have not been fully elucidated. However, putative functions have been proposed for several, based on their sites of expression and/or on the activity of orthologous proteins.

For instance, accumulating data suggest that several kallikreins may be implicated in the processing of peptide hormones in the endocrine pancreas. Immunohistochemical studies indicate that hK1, 6, 10, and 13 are all strongly expressed in the islets of Langerhans, within the specialized β, α, δ and pancreatic polypeptide cells that synthesize insulin, glucagon, somatostatic, and pancreatic polypeptide, respectively (120, 121, 127, 167, 206-208). Accordingly, these kallikreins may participate in prohormone activation, possibly in cooperation with other prohormone convertases, such as PC1 and PC2, which also colocalize with kallikreins in similar cell types of the endocrine pancreas (208). In fact, the activation of proinsulin to mature insulin by hK1 has already been documented (209).

Several suggest that kallikreins play a role in the normal physiology of the skin, particularly in epidermal homeostasis. Both hK5 (47) and hK7 (28) have been isolated and cloned from the stratum corneum, the outermost layer of the skin. They are proposed to function in the degradation of intercellular structures (131, 210), such as desmosomes (211-214), connecting the corneocytes, thereby decreasing cellular cohesiveness and facilitating cell shedding or desquamation during the terminal stages of epidermal turnover. However, because hK5 and hK7 preferentially cleave only a subset of desmosomal proteins (214), additional trypsin and chymotrypsin-like proteases are implicated in stratum corneum desquamation (215), including many other members of the kallikrein family (108). In addition to desquamation, hK7 may also play a role in skin pathophysiology, including pathologic keratinization (130), psoriasis (216), and in inflammatory reactions, due to its ability to activate proinflammatory cytokines, such as interleukin-1β (IL-1β; refs. 217, 218).

We have recently published a review discussing the potential roles of kallikreins in the central nervous system (219). Putative functions for hK6 and hK8 have been extrapolated from the experimentally verified actions of their rodent orthologs. Given the high amino acid sequence identity among hK6 and hK8 and their rodent orthologs (∼70%), it is conceivable that the proteins exhibit similar activities. For example, the rat ortholog of hK6, called myelencephalon-specific protease (MSP), may play a role in the regulation of central nervous system demyelinating disease (220-223), including the development of multiple sclerosis lesions (221), while the mouse ortholog may function in myelination and myelin turnover (224). As well, human kallikrein 6 is implicated in the development of Alzheimer's disease partly due to its ability to cleave amyloid precursor protein, in vitro, and possibly generate β-amyloid peptides (31, 178), which are known to aggregate and form one of the major pathologic lesions characteristic of this disease. Several reports indicate that mouse hK8/neuropsin might be involved in synaptogenesis, neural development (225), regulation of long-term potentiation (LTP; refs. 226, 227) and seizures in kindled brain (228).

Furthermore, based on the expression patterns and suggested roles of its mouse and porcine orthologs, the human hK4 protein may likely be involved in tooth development via enamel matrix protein degradation and/or processing during dental enamel formation (229-232).

Moreover, in addition to hK3 (233) and hK6 (69, 178), our preliminary data suggest that several other kallikreins, including hK5, hK13, and hK14, are able to cleave components of the ECM in vitro. Hence, kallikreins may also function in tissue remodeling, similar to matrix metalloproteases (234).

Circumstantial evidence suggests that cross-talk likely exists among members of the human kallikrein gene family and with proteases of other catalytic classes. On the basis of the colocalization of kallikrein genes to the same chromosomal locus; their coordinated regulation by steroid hormones, coexpression in tissues, and biological fluids; and the reported ability of certain kallikreins to autoactivate and potentially activate other kallikreins and proteases, it has been presumed that this family may participate in a proteolytic cascade pathway (235). The best examples of well-established enzymatic cascades involving serine proteases include the blood coagulation, fibrinolytic, and digestive cascades (236, 237). These cascades are characterized by a series of zymogen or pro-enzyme activations, in which the activated form of one enzyme catalyzes the activation of the following zymogen, and by the rapid amplification of the initial signal during their progression.

As illustrated in Fig. 1, all 15 kallikrein genes colocalize to 19q13.4. The colocalization of genes encoding proteins that take part in the same pathway is not uncommon in the human genome. For example, several serine proteases involved in sequential steps of the coagulation cascade are encoded by tandemly colocalized genes and some may share a common ancestor, similar to the kallikrein family (236, 238, 239). Kallikrein genes are also coordinately regulated by steroid hormones and coexpressed in similar tissues (e.g., skin, prostate, breast, pancreas) and found in biological fluids (e.g., seminal plasma, milk of lactating women) under normal conditions, as discussed above. The parallel pattern of differential kallikrein expression in malignancy, such as the concurrent up-regulation of KLK5, KLK6, KLK7, KLK8, KLK10, KLK11, and KLK14 in ovarian cancer (240), further substantiates the possible existence of a steroid hormone–driven cascade.

As shown in Table 2, all hK pro-peptide cleavage sites contain a P1 Arg or Lys residue, with the exception of hK4, indicating that pro-hKs generally require the activity of a trypsin-like serine protease for activation. For instance, it has been shown that trypsin can convert pro-hK5, pro-hK6, pro-hK7, and pro-hK15 into their active forms, while enterokinase can activate pro-hK11 (28, 36, 47, 64, 167). Because most kallikreins possess trypsin-like specificity, many are implicated in autoactivation and the activation of other pro-hKs as well. For instance, hK2 (174, 201), hK6 (31, 178), and hK13 (179) are all capable of autoactivation and may, therefore, be involved in the initiation and maintenance of a cascade, similar to factor XI of the intrinsic coagulation pathway (241). As well, experimental evidence has shown that recombinant hK2, hK4, and hK15 can readily activate pro-PSA, in vitro (53, 173, 202, 203, 242). Brattsand and Egelrud (47) also hypothesize that hK5 may potentially activate pro-hK7 in the skin.

Kallikreins may also be implicated in additional pathways involving proteases of similar or different catalytic types. This is evident from the reported ability of hK2 and hK4 to activate the pro-form of uPA (242, 243), a serine protease that converts the serine protease, plasminogen to plasmin, which in turn degrades the ECM and activates members of the matrix metalloprotease family (244). In addition to plasmin, kallikreins may also be involved in the activation of pro-matrix metalloproteases, because matrix metalloproteases require the activity of trypsin-like serine proteases for cleavage of their pro-peptides (245, 246). Interestingly, porcine hK1 was found to activate type IV collagenase, a matrix metalloprotease family member (247). These findings implicate kallikreins in the promotion of tumor invasion and metastasis (further discussed below). Conversely, some members of the kallikrein family, namely hK3 (248), hK13 (179), and hK67

7

G. Sotiropoulou, personal communication.

are able to cleave plasminogen, causing the release of biologically active angiostatin-like fragments known to inhibit angiogenesis (249).

Protease-activated receptors (PAR) comprise a small subfamily of G protein–coupled receptors, through which serine proteases mediate their hormone-like effects on cells (250). Unlike most receptors, PARs are stimulated by serine protease cleavage of an extracellular NH2-terminal segment, generating a new NH2-terminal sequence that acts as a tethered ligand and interacts with the second extracellular loop (251). Thus, the protease changes the conformational structure of the receptor, such that it acts as its own activator. Due to the coupling of PARs with several G protein family members, a complex network of intracellular signaling pathways may be activated, leading to changes in morphology, proliferation, survival, cell mobility, and gene transcription, shown to be important in the physiology of the vascular and nervous systems (252-254). Out of the four human PARs identified to date, PAR-1 (250), -3 (255), and -4 (256) are relatively specific for thrombin, whereas PAR-2 is not activated by thrombin and seems to have a broader range of cognate proteases, including trypsin, tryptase, and coagulation factors VIIa and Xa (257, 258). However, PAR-1 does not restrict activation by other proteases, because it may also be activated by coagulation factor Xa (259), plasmin (260), and the anticoagulant protein C (261). The absence of the unique thrombin-complementary extracellular domain from PAR-2 and PAR-4 leads to the concept of “generic PARs” and renders them as candidates for other serine proteases, including human kallikreins. Recently, a trypsin-like serine protease named P22, enzymatically similar to rat hK8, was isolated from rat brain and shown to activate PAR-2 (262). The potential involvement of human kallikreins in the activation of PARs should be explored.

Carcinogenesis is a complex process that involves alterations at the DNA, mRNA, and protein levels. The main goal of cancer research is to identify these alterations and determine their effects on the tumor phenotype. Accumulating reports indicate that the human kallikrein family is implicated in cancer. All 15 kallikrein genes are differentially expressed in cancer, primarily in hormone-related malignancies, at the mRNA and/or protein levels. For instance, numerous studies have shown that kallikreins 4, 5, 6, 7, 8, 10, 11, 13, 14, and 15 are overexpressed in ovarian carcinoma tissues, serum, and/or cell lines at the mRNA and/or protein levels (30, 34, 71, 110, 115, 116, 119, 123, 263-272). The up-regulation of kallikreins 5, 6, 7, 8, 10, 11, and 14 in ovarian cancer was further verified in silico via digital differential display and X-profiler analyses of kallikrein gene expression in normal and cancerous ovarian tissues and cell lines by Yousef et al. (240). In contrast to ovarian cancer, kallikrein genes 3, 10, 12, 13, and 14, are down-regulated in breast cancer tissues and/or cell lines at the mRNA level (29, 49, 50, 52, 125, 273-276), while the KLK6 gene is down-regulated in metastatic breast cancer sites and up-regulated in primary breast tumors (30). In silico analyses of kallikrein mRNA expression levels in normal and cancerous breast tissues and cell lines suggests that at least four kallikrein genes, namely KLK5, 6, 8, and 10 are down-regulated in breast cancer (277), partially consistent with the previous findings. Although human kallikrein 5 and 14 mRNA levels are reduced in breast cancer, elevated serum levels of the hK5 and 14 proteins were observed in a subgroup of breast cancer patients (110, 116). This discrepancy between kallikrein 5 and 14 mRNA and serum protein levels in breast cancer has also been observed for hK3/PSA in prostate cancer. In these cases, the elevation of hK proteins in the serum may be due to angiogenesis and/or destruction of glandular architecture during carcinogenesis, thereby facilitating the outflow of hKs into the circulation. With respect to prostate cancer, KLK2, KLK3, KLK5, KLK6, KLK10, and KLK13 are down-regulated compared with normal adjacent tissue (117, 122, 278-282), whereas KLK11, KLK14, and KLK15 are overexpressed (115, 283, 284). Additionally, the expression of KLK5, KLK10, and KLK14 is markedly reduced in cancerous versus normal testicular tissues at the mRNA level (52, 285, 286). (For a recent review describing the association of kallikreins with testicular cancer, see ref 287.)

In addition to hormone-related cancers discussed above, kallikrein expression is dysregulated in several other malignancies. A recent microarray analysis profiling the gene expression patterns in human lung adenocarcinomas indicated that KLK11 is uniquely overexpressed in a subgroup of neuroendocrine carcinomas (288). Another microarray study has characterized differential transcription profiles in pancreatic ductal adenocarcinomas and showed that KLK10 is one of the most highly and specifically overexpressed genes in pancreatic cancer compared with normal and benign pancreas tissues (289). Furthermore, the KLK10 gene is down-regulated in acute lymphblastic leukemia (168).

Emerging data indicate that many alternative kallikrein transcripts are also differentially expressed in cancer and some are even cancer-specific. Dong et al. (101) have recently documented the overexpression of a KLK5 variant with a short 5′ UTR and a KLK7 variant with a longer 3′ UTR in ovarian cancer cell lines, compared with normal ovarian epithelial cells. A KLK5 splice variant, denoted KLK5 splice variant 1, is up-regulated in ovarian cancer tissues, but down-regulated in prostate cancer tissues compared with normal.8

8

Our data submitted for publication.

Two novel mRNA splice variants of the KLK8 gene, missing either two or three coding exons, are overexpressed at relatively high levels in cancerous ovarian tissues, compared with normal ovarian tissues, in which they were not detected (267). The KLK11 gene has two transcript variants named the brain type and prostate type (36), both of which are overexpressed in cancerous prostate versus normal tissues (290). Moreover, the work of Chang et al. (133) has revealed that the KLK13 gene possesses at least five tissue-specific splice variants expressed exclusively the testis, in contrast to classical KLK13 mRNA which is predominately expressed in a variety of tissues including the breast, prostate, testis, and salivary gland (49). These KLK13 splice variants are expressed in a fraction of morphologically normal testicular tissues, but absent in the adjacent cancerous tissues (133).

The mechanisms giving rise to the differential expression of kallikrein genes in cancer have not been fully elucidated. However, epigenetic modifications, specifically the hypermethylation of coding exon 3, are responsible for the down-regulation of KLK10 in breast, prostate, and ovarian cancers and acute lymphoblastic leukemia (refs. 167, 168) and our data submitted for publication). This epigenetic mechanism is an important cause of gene silencing in carcinogenesis (291). Also, considering that steroid hormones are implicated in the etiology of hormone-related malignancies, such as ovarian, breast, prostate, and testicular cancers (138), and are also known to regulate kallikrein gene expression, it may be possible that kallikreins are part of a steroid hormone–driven (cascade) pathway that is activated during the promotion and progression of cancer.

With respect to their involvement in the pathogenesis of cancer, kallikreins seem to have a dual role, because they can either promote or inhibit carcinogenesis. For one, many kallikreins are directly and/or indirectly involved in the degradation of ECM proteins, which facilitates tumor invasion and metastasis. As listed in Table 2, hK2, 3, 5, 6, 13, and 14 can directly catalyze the hydrolysis of several ECM proteins (refs. 69, 178, 204, 205, 233 and our unpublished data). hK2 and hK4 are indirectly involved via activation of uPA, leading to plasminogen activation, ultimately resulting in ECM degradation (242, 243). Furthermore, a synthetic hK1 inhibitor was recently found to suppress the invasiveness in human breast cancer cell lines by 33% in matrigel invasion assays (292), suggesting that it has a role in facilitating tumor cell migration, via ECM cleavage. These findings are further supported by the numerous reports indicating that kallikrein overexpression is associated with poor prognosis in cancer patients (Tables 4Table 5-6). As well, hK3 can also cleave insulin-like growth factor binding protein 3 (IGFBF-3), thus, liberating insulin-like growth factor, which is a mitogen for prostatic stromal and epithelial cells (293) and activate transforming growth factor β , thereby stimulating cell detachment and facilitating tumor spread (294). hK1 is present in colon, breast, lung, stomach, pituitary, uterine, and esophageal cancer cells and may be involved in malignant transformation by stimulating proliferation of tumor cells and increasing vascular permeability (148, 198, 295-302). hK1, via kinin action, also enhances vascularity, mitogenicity, metastasis, and regulates angiogenesis (303-305).

Kallikreins may also inhibit carcinogenesis. KLK10 is thought to be a tumor suppressor gene, by virtue of its down-regulation in several cancers (29, 125, 168, 276) and because transfection of this gene into the tumorigenic breast cancer cell line MDA-MB-231 reduced its anchorage-independent growth and the tumor formation of nude mice inoculated with this KLK10-transfected cell line was significantly reduced. hK3 may also act as a tumor suppressor, an inducer of apoptosis (306), and as a negative regulator of breast cancer cell growth (307). hK3, 6, and 13 are also implicated in the inhibition of angiogenesis, via their release of angiostatin-like fragments from plasminogen (refs. 179, 248, 308 and G. Sotiropoulou, personal communication). These studies may help to explain, in part, why certain kallikreins are markers of favorable prognosis for cancer patients (Tables 4,Table 5,-6).

Cancer biomarkers are fundamental tools that aid in evaluating cancer risk, screening, diagnosis, clinical staging, estimating tumor volume, monitoring, assessing prognosis, evaluating success of treatment, detecting disease recurrence, and predicting a likely response to therapy to improve patient management and outcomes (309). Among all biomarkers to date, hK3/PSA has had the greatest impact in clinical practice, for the screening, diagnosis, staging, and monitoring of prostate cancer (310). Most other biomarkers often lack the desired sensitivity and specificity and, thus, the search for more informative markers continues.

The tissue kallikrein family has proved to be a rich source of cancer biomarkers (311-313). In addition to hK3, many other kallikreins exhibit altered mRNA and/or protein expression levels within the tissues and/or serum of cancer patients and represent prospective biomarkers for early detection, prognosis, or monitoring of certain hormone-dependent malignancies, as summarized in Tables 4,Table 5,-6. For instance, serum hK2 may function as an alternate or complementary biomarker to hK3 for prostatic diseases (314, 315). Importantly, this kallikrein may aid in the differential diagnosis between prostate cancer and benign prostatic hyperplasia (316) as well as the identification of organ-confined versus non-organ–confined disease (317). With the recent developments of highly sensitive and specific immunoassays for hK5, hK6, hK8, hK10, hK11, and hK14 proteins, elevated levels of these kallikreins were observed in the tissues and/or serum of a proportion of ovarian cancer patients (110, 113, 115, 116, 266, 270, 271, 318). Pre-surgical serum hK6 and hK10 levels increase the diagnostic sensitivity of CA125 in patients with early stage (I/II) ovarian cancer and are associated with poor patient prognosis. Thus, serum hK6 and hK10 may complement CA125 for early detection of ovarian cancer. Serum hK5 and hK14 levels are also increased in ∼40% of women with breast cancer, whereas serum hK11 is elevated in 60% of men with prostate cancer. The latest study by Nakamura et al. (319) also shows that serum hK11 levels and the hK11/total PSA ratio are both significantly lower in patients with prostate cancer than in BPH patients, suggesting that serum hK11/total PSA may aid in differential diagnosis. Using the hK11/total PSA ratio at 90% sensitivity, it would be possible to avoid ∼50% unnecessary prostatic biopsies not evaded by the %free-PSA test (%free PSA <20). Furthermore, complexed forms of hK2 and hK3 with various plasma inhibitors (320), degraded forms of hK2 and hK3 (321), and the altered glycosylation patterns observed between hK3 produced by normal and cancerous prostate tissues (322), may also have clinical utility in prostate cancer diagnostics and prognostics.

In addition to their clinical value as serologic biomarkers, kallikrein mRNA and protein levels within cancerous tissues are often associated with patient prognosis. As listed in Table 4, KLK4, KLK5/hK5, KLK6/hK6, KLK7, hK10, and KLK15 are markers of poor prognosis in ovarian cancer (263-265, 269, 271, 272, 323, 324). That is, higher kallikrein mRNA and/or protein levels were found to correlate with more aggressive forms of this disease and a decreased disease-free and overall survival. Conversely, the remaining subset of kallikreins, namely KLK8, KLK9, hK11, hK13, and KLK14, are markers of favorable prognosis (124, 147, 267, 325, 326). Higher levels of their mRNA or protein levels predominate in earlier stage disease and are associated with increased disease-free and overall survival. The expression of these kallikreins in ovarian cancer may also be clinically useful in determining the prognosis in subgroups of patients. For instance, a subgroup of kallikreins (kallikreins 4, 6, and 10) are highly expressed in serous epithelial ovarian tumors, whereas higher expression of another group (kallikreins 5, 11, and 13) is more frequently found in non-serous tumors. These data suggest that certain kallikreins may be used as determinants of prognosis in the subgroups of ovarian cancer patients stratified by histotype as well.

Table 4.

Human Kallikreins as Ovarian Cancer Biomarkers (Messenger RNA and/or Protein Level)

Kallikrein Gene (KLK)/Protein (hK)Samples UsedClinical ApplicationsReferences
KLK4 mRNA from normal and cancerous ovarian tissues Unfavorable prognostic marker (119, 263) 
KLK5 mRNA from normal and cancerous ovarian tissues Unfavorable prognostic marker (264) 
KLK5/hK5 mRNA and cytosolic extracts from normal and cancerous ovarian tissues Unfavorable prognostic marker (101) 
hK5 ovarian cancer cytosols Unfavorable prognostic marker our data submitted for publication 
 serum and tissue Marker of diagnosis (110) 
KLK6/hK6 mRNA and extracts from normal, benign, and cancerous ovarian tissues Unfavorable prognostic marker (265) 
hK6 ovarian cancer cytosols Unfavorable prognostic marker (323) 
 serum Marker of diagnosis, prognosis, and monitoring (266, 318) 
KLK7 mRNA from cancerous ovarian tissue Unfavorable prognostic marker (324) 
KLK7/hK7 mRNA and extracts from normal and cancerous ovarian tissues Unfavorable prognostic marker (101, 123) 
KLK8 mRNA from ovarian cancer tissues Favorable prognostic marker (267) 
hK8 serum and tissue Marker of diagnosis, prognosis, and monitoring (268) 
KLK9 mRNA from ovarian cancer tissues Favorable prognostic marker (124) 
hK10 serum and tissue Marker of diagnosis, prognosis, and monitoring (270, 271) 
 normal, benign, and cancerous ovarian cytosols Unfavorable prognostic marker (269) 
hK11 ovarian cancer cytosols Favorable prognostic marker (325) 
 serum Marker of diagnosis (115) 
hK13 ovarian cancer cytosols Favorable prognostic marker (326) 
KLK14 mRNA from normal, benign, and cancerous ovarian tissues Favorable prognostic marker (147) 
hK14 serum and tissue Marker of diagnosis (116) 
KLK15 mRNA from benign and cancerous ovarian tissues Unfavorable prognostic marker (272) 
Kallikrein Gene (KLK)/Protein (hK)Samples UsedClinical ApplicationsReferences
KLK4 mRNA from normal and cancerous ovarian tissues Unfavorable prognostic marker (119, 263) 
KLK5 mRNA from normal and cancerous ovarian tissues Unfavorable prognostic marker (264) 
KLK5/hK5 mRNA and cytosolic extracts from normal and cancerous ovarian tissues Unfavorable prognostic marker (101) 
hK5 ovarian cancer cytosols Unfavorable prognostic marker our data submitted for publication 
 serum and tissue Marker of diagnosis (110) 
KLK6/hK6 mRNA and extracts from normal, benign, and cancerous ovarian tissues Unfavorable prognostic marker (265) 
hK6 ovarian cancer cytosols Unfavorable prognostic marker (323) 
 serum Marker of diagnosis, prognosis, and monitoring (266, 318) 
KLK7 mRNA from cancerous ovarian tissue Unfavorable prognostic marker (324) 
KLK7/hK7 mRNA and extracts from normal and cancerous ovarian tissues Unfavorable prognostic marker (101, 123) 
KLK8 mRNA from ovarian cancer tissues Favorable prognostic marker (267) 
hK8 serum and tissue Marker of diagnosis, prognosis, and monitoring (268) 
KLK9 mRNA from ovarian cancer tissues Favorable prognostic marker (124) 
hK10 serum and tissue Marker of diagnosis, prognosis, and monitoring (270, 271) 
 normal, benign, and cancerous ovarian cytosols Unfavorable prognostic marker (269) 
hK11 ovarian cancer cytosols Favorable prognostic marker (325) 
 serum Marker of diagnosis (115) 
hK13 ovarian cancer cytosols Favorable prognostic marker (326) 
KLK14 mRNA from normal, benign, and cancerous ovarian tissues Favorable prognostic marker (147) 
hK14 serum and tissue Marker of diagnosis (116) 
KLK15 mRNA from benign and cancerous ovarian tissues Unfavorable prognostic marker (272) 

With respect to breast cancer (Table 5), the mRNA expression levels of KLK5, KLK7, and KLK14 in breast tumors are indicative of a poor patient prognosis (327-329), while higher levels of KLK9, KLK13, and KLK15 mRNA and the hK3 protein forecast a favorable disease outcome (158, 275, 330, 331). Furthermore, high levels of hK3 and hK10 proteins in breast carcinomas are significantly related to a poor response to tamoxifen therapy (332, 333).

Table 5.

Human Kallikreins as Breast Cancer Biomarkers (Messenger RNA and/or Protein Level)

Kallikrein Gene (KLK)/Protein (hK)Samples UsedClinical ApplicationsReferences
hK3 serum and tissue Marker of diagnosis and prognosis reviewed in ref. 402 
KLK5 mRNA from breast cancer tissues Unfavorable prognostic marker (158) 
hK5 serum Diagnostic marker (110) 
KLK7 mRNA from breast cancer tissues Unfavorable prognostic marker (328) 
KLK9 mRNA from breast cancer tissues Favorable prognostic marker (330) 
hK10 breast cancer cytosols Predictive value (333) 
KLK13 mRNA from breast cancer tissues Favorable prognostic marker (331) 
KLK14 mRNA from breast cancer tissues Unfavorable prognostic marker (329) 
hK14 serum and tissue Diagnostic marker (116) 
KLK15 mRNA from breast cancer tissues Favorable prognostic marker (158) 
Kallikrein Gene (KLK)/Protein (hK)Samples UsedClinical ApplicationsReferences
hK3 serum and tissue Marker of diagnosis and prognosis reviewed in ref. 402 
KLK5 mRNA from breast cancer tissues Unfavorable prognostic marker (158) 
hK5 serum Diagnostic marker (110) 
KLK7 mRNA from breast cancer tissues Unfavorable prognostic marker (328) 
KLK9 mRNA from breast cancer tissues Favorable prognostic marker (330) 
hK10 breast cancer cytosols Predictive value (333) 
KLK13 mRNA from breast cancer tissues Favorable prognostic marker (331) 
KLK14 mRNA from breast cancer tissues Unfavorable prognostic marker (329) 
hK14 serum and tissue Diagnostic marker (116) 
KLK15 mRNA from breast cancer tissues Favorable prognostic marker (158) 

Several kallikreins also have prognostic/predictive value in prostate carcinoma (Table 6). For example, lower tissue hK3 concentration is associated with more aggressive forms of this cancer, such that tumors expressing high levels are associated with a favorable prognosis (279, 334). Higher KLK5 and KLK11 mRNA levels also indicate a favorable prognosis (281, 290). Moreover, hypermethylation of the KLK10 gene in coding exon 3 is an independent prognostic marker of decreased disease-free survival in children and adults and in the separate analysis of adults with acute lymphoblastic leukemia (168).

Table 6.

Human Kallikreins as Prostate Cancer Biomarkers (Messenger RNA and/or Protein Level)

Kallikrein Gene (KLK)/Protein (hK)Samples UsedClinical ApplicationsReferences
hK2 serum and tissue Marker of diagnosis, prognosis, and monitoring (315, 321, 403) 
hK3 serum and tissue Marker of diagnosis, prognosis, and monitoring (321) 
KLK5 mRNA from matched normal and prostate cancer tissues Favorable prognostic marker (281) 
KLK11 mRNA from matched normal and prostate cancer tissues Favorable prognostic marker (290) 
hK11 serum Diagnostic marker (319) 
Kallikrein Gene (KLK)/Protein (hK)Samples UsedClinical ApplicationsReferences
hK2 serum and tissue Marker of diagnosis, prognosis, and monitoring (315, 321, 403) 
hK3 serum and tissue Marker of diagnosis, prognosis, and monitoring (321) 
KLK5 mRNA from matched normal and prostate cancer tissues Favorable prognostic marker (281) 
KLK11 mRNA from matched normal and prostate cancer tissues Favorable prognostic marker (290) 
hK11 serum Diagnostic marker (319) 

The discovery of cancer-specific mRNA transcript variants that occur exclusively or with higher frequency in cancer cells and which are detectable by biopsy or in body fluids, may serve as useful diagnostic biomarkers. The most well-characterized cancer-specific splice variant biomarkers include those of the CD44 and Wilm's tumor (WT1) genes (335, 336). As previously discussed, preliminary studies indicate that variant kallikrein transcripts are differentially expressed and/or expressed specifically in cancer and may, therefore, constitute a new generation of cancer biomarkers within the kallikrein family.

For instance, Slawin et al. (337) have recently developed a preoperative KLK2 splice variant–specific reverse transcription-PCR that is useful for detecting prostate cancer metastasis and helps predict pathologic lymph node positivity in men with clinically localized prostate cancer. Tanaka et al. (105) have reported the existence of an alternatively spliced form of the KLK3 gene that is expressed in 13 of 18 (72.2%) noncancerous and 4 of 5 (80.0%) cancerous prostate tissues, but in only 3 of 12 (25.0%) blood samples from prostate cancer patients. The difference in KLK3 variant expression levels between noncancerous prostate tissues versus blood samples from cancer patients was statistically significant (P = 0.011). David et al. (132) have reported the identification of two splice variants of the KLK2 and KLK3 genes that result from inclusion of intronic sequences adjacent to the first exon, denoted K-LM and PSA-LM, respectively. With the exception of the signal peptide, K-LM and PSA-LM transcripts encode protein isoforms that are entirely different than the classical hK2 and hK3 proteins. As such, polyclonal antibodies were generated against synthetic peptides derived from amino acid sequences unique to each variant protein. Immunohistochemistry of prostate sections using these polyclonal antibodies indicated that the K-LM and PSA-LM proteins are detected only in the secreting cells of the tubule lumen and Western blot analysis indicated that the K-LM protein is present in seminal plasma, similar to the classical forms of hK2 and hK3. Furthermore, a recent study indicates that KLK3 may actually produce at least 15 transcripts, which can encode eight putative protein isoforms (97). Reverse transcription-PCR analysis indicates that at least five splicing isoforms are expressed in normal, benign prostatic hyperplastic, and cancerous tissues. Collectively, these KLK2 and KLK3 variants may supplement hK3/PSA diagnostics. Using quantitative reverse transcription-PCR, Nakamura et al. (290) compared the expression of the prostate and brain-type KLK11 isoforms in matched normal and cancerous prostatic tissues. Both variants were overexpressed in cancerous prostate versus normal tissues and lower expression of prostate-type KLK11 was associated with higher tumor stage, grade and Gleason score. No such association was seen with the brain-type isoform. These data suggest that KLK11 splice variants may have clinical value as biomarkers for prostate cancer diagnosis and prognosis. Variant transcripts of KLK5, 8, and 13 are also differentially expressed in cancer, as discussed in a previous section. The biological and clinical significance of these variant kallikrein transcripts/proteins remains to be elucidated.

Recently, it has become possible to combine the diagnostic, prognostic, and predictive value of multiple biomarkers into models, which have the ability to discriminate better than single biomarkers alone (338-342). For example, the use of logistic regression, decision tress, discriminant analysis, and artificial neural networks can outperform single biomarker analysis in diagnostic, prognostic, and predictive applications. Therefore, the combination of a subset of classical and/or variant kallikreins into a multiparametric panel may provide superior diagnostic/prognostic information than that of the single analytes alone. Further studies are warranted to evaluate this hypothesis.

Single-nucleotide polymorphism (SNP) within candidate genes can affect coding sequences, transcriptional regulation, and splicing and may confer increased susceptibility or resistance to complex diseases, such as cancer. As such, SNPs are considered potential markers of cancer risk and progression and can help to define resistance to therapeutic regimens. Several SNPs have been reported in the human kallikrein locus, within KLK1 (343), KLK2 (344), and KLK10 (345) genes and the KLK3 (155, 346, 347) promoter region. The KLK2 SNP (C→T) in exon 5 changes the amino acid from Arg226 to Trp266, leading to an active (C allele; Arg226) and inactive (T allele; Trp266) form of hK2 (344). A recent study has found a strong positive relationship between this polymorphism with serum hK2 levels and prostate cancer risk, that is, the T allele is associated with lower hK2 levels, but a higher risk of cancer (348). Regarding KLK3, three SNPs are in the proximal promoter at positions −158, −205, and −252, which may be implicated in breast and/or prostate cancer susceptibility. For example, concerning the SNP at position −158 (G→A), individuals homozygous for the G allele showed higher hK3 tumor concentrations and an increased overall survival than those homozygous for the A allele (349). Depending on the study, either the G or A allele of SNP −158 was also shown to be associated with the risk of advanced prostate cancer or an earlier onset of prostate cancer in Caucasian men (350-352), an association not found in a study involving Japanese men (353). The polymorphisms at positions −205 and −252 may be associated with mRNA expression levels of KLK3 (354), whereas the −252 SNP was not linked to prostate cancer risk and progression in two separate Japanese studies (353, 355). Lastly, Bharaj et al. (345) have identified a few of SNPs within exons 3 and 4 and intron 5 of the KLK10 gene. The most significant SNP is in codon 50 within exon 3 (C→T) and changes the amino acid in this position from Ala to Ser. The prevalence of the T allele was significantly higher in prostate cancer patients in comparison to control subjects, and may, therefore, be associated with prostate cancer risk (345).

Kallikreins may also constitute potential drug targets, therapeutic agents, and candidates for passive or active immunotherapy, once their biological pathways are delineated. For instance, the identification of CD4 positive T cells specific for naturally processed KLK4-derived epitopes within the T-cell repertoire of normal males support the use of KLK4 as a target for whole gene-, protein-, or peptide-based vaccine strategies against prostate cancer (356). This kallikrein may also represent a target for immunotherapy because anti-hK4 antibodies were only present in the serum of males with prostate cancer (357).

With the discovery of the complete hK family, comprising a total of 15 serine protease genes on 19q13.4, the genomic era of kallikrein research is nearing its end. On the basis of tissue expression patterns and putative substrates, tissue kallikreins are implicated in diverse physiologic processes, from the regulation of cell growth to tissue remodeling, where they may act individually or in cascade pathway(s). Countless reports have also indicated an association between dysregulated kallikrein expression and cancer and the carcinogenic process and the potential use of kallikreins as diagnostic/prognostic biomarkers for cancer. However, many questions remain unanswered with respect to the exact role of many tissue kallikreins in normal and pathophysiology. The future must encompass the identification of physiologic substrates, delineation of the functional intersections between kallikreins and other proteolytic systems including those involved in cell signaling, a better understanding of modes of regulation, and unveiling the relevance of the complete kallikrein transcriptome and proteome including variant mRNA transcripts and proteins. With respect to cancer biomarker and drug discovery, the tissue kallikrein family is a gold mine waiting to be fully unearthed. Therefore, another important goal for the future will involve further defining the clinical utility of kallikreins as biomarkers for cancer, as single analytes or in combination with a number of suitable molecules in a multiparametric model. Thus, the post-genomic era poses a new set of challenges for research in this subgroup of the human degradome.

1
Nomenclature Committee of the International Union of Biochemistry and Molecular Biology (NC-IUBMB) Enzyme Nomenclature 1992. Recommendations of the Nomenclature Committee of the International Union of Biochemistry and Molecular Biology on the Nomenclature and Classification of Enzymes. Orlando, FL: Academic Press; 1992.
2
Rawlings ND, Barrett AJ. MEROPS: the peptidase database.
Nucleic Acids Res
1999
;
27
:
325
-31.
3
Lopez-Otin C, Overall CM. Protease degradomics: a new challenge for proteomics.
Nat Rev Mol Cell Biol
2002
;
3
:
509
-19.
4
Puente XS, Sanchez LM, Overall CM, Lopez-Otin C. Human and mouse proteases: a comparative genomic approach.
Nat Rev Genet
2003
;
4
:
544
-58.
5
Barrett AJ, Rawlings ND, Woessner JF. Handbook of proteolytic enzymes. London: Academic Press; 1998.
6
Rawlings ND, Tolle DP, Barrett AJ. MEROPS: the peptidase database.
Nucleic Acids Res
2004
;
32 Database issue
:
D160
-4.
7
Neurath H. Proteolytic enzymes, past and present.
Fed Proc
1985
;
44
:
2907
-13.
8
Perona JJ, Craik CS. Structural basis of substrate specificity in the serine proteases.
Protein Sci
1995
;
4
:
337
-60.
9
Sim RB, Laich A. Serine proteases of the complement system.
Biochem Soc Trans
2000
;
28
:
545
-50.
10
Del Rosso M, Fibbi G, Pucci M, et al. Multiple pathways of cell invasion are regulated by multiple families of serine proteases.
Clin & Exp Metastasis
2000
;
19
:
193
-207.
11
Noel A, Gilles C, Bajou K, et al. Emerging roles for proteinases in cancer.
Invasion Metastasis
1997
;
17
:
221
-39.
12
Turgeon VL, Houenou LJ. The role of thrombin-like (serine) proteases in the development, plasticity and pathology of the nervous system.
Brain Res Brain Res Rev
1997
;
25
:
85
-95.
13
Meyer W. Distribution of tissue kallikrein in the integumental blood vessel system of wild mammals.
Cell Mol Biol (Noisy-le-grand)
2001
;
47 Online Pub
:
125
-30.
14
Yousef GM, Diamandis EP. An overview of the kallikrein gene families in humans and other species: emerging candidate tumour markers.
Clin Biochem
2003
;
36
:
443
-52.
15
Olsson A, Lilja H, Lundwall A. Taxon-specific evolution of glandular kallikrein genes, and identification of a progenitor of prostate-specific antigen. Genomics. In press 2004.
16
Chapdelaine P, Gauthier E, Ho-Kim MA, Bissonnette L, Tremblay RR, Dube JY. Characterization and expression of the prostatic arginine esterase gene, a canine glandular kallikrein.
DNA Cell Biol
1991
;
10
:
49
-59.
17
Olsson A, Lundwall A. Organization and evolution of the glandular kallikrein locus in Mus musculus.
Biochem Biophys Res Commun
2002
;
299
:
305
-11.
18
Yousef GM, Diamandis EP. The new human tissue kallikrein gene family: structure, function, and association to disease.
Endocr Rev
2001
;
22
:
184
-204.
19
Yousef GM, Kopolovic AD, Elliott MB, Diamandis EP. Genomic overview of serine proteases.
Biochem Biophys Res Commun
2003
;
305
:
28
-36.
20
Frey EK. Zusammenhange zwischen Herzabeit und Nierentatigkeit.
Arch Klin Chir
1926
;
142
:
663
-9.
21
Frey EK, Kraut H. Ein neues Kreislaufhormon und seine Wirkung.
Arch Exp Pathol Pharmakol
1928
;
133
:
1
-56.
22
Kraut H, Frey EK, Werle E. Der Nachweis eines Kreislaufhormon in de pankreasdruse.
Hoppe-Seyler Z Physiol Chem
1930
;
189
:
97
-106.
23
Baker AR, Shine J. Human kidney kallikrein: cDNA cloning and sequence analysis.
DNA
1985
;
4
:
445
-50.
24
Fukushima D, Kitamura N, Nakanishi S. Nucleotide sequence of cloned cDNA for human pancreatic kallikrein.
Biochemistry
1985
;
24
:
8037
-43.
25
Schedlich LJ, Bennetts BH, Morris BJ. Primary structure of a human glandular kallikrein gene.
DNA
1987
;
6
:
429
-37.
26
Riegman PH, Vlietstra RJ, van der Korput JA, Romijn JC, Trapman J. Characterization of the prostate-specific antigen gene: a novel human kallikrein-like gene.
Biochem Biophys Res Commun
1989
;
159
:
95
-102.
27
Riegman PH, Vlietstra RJ, Klaassen P, et al. The prostate-specific antigen gene and the human glandular kallikrein-1 gene are tandemly located on chromosome 19.
FEBS Lett
1989
;
247
:
123
-6.
28
Hansson L, Stromqvist M, Backman A, Wallbrandt P, Carlstein A, Egelrud T. Cloning, expression, and characterization of stratum corneum chymotryptic enzyme. A skin-specific human serine proteinase.
J Biol Chem
1994
;
269
:
19420
-6.
29
Liu XL, Wazer DE, Watanabe K, Band V. Identification of a novel serine protease-like gene, the expression of which is down-regulated during breast cancer progression.
Cancer Res
1996
;
56
:
3371
-9.
30
Anisowicz A, Sotiropoulou G, Stenman G, Mok SC, Sager R. A novel protease homolog differentially expressed in breast and ovarian cancer.
Mol Med
1996
;
2
:
624
-36.
31
Little SP, Dixon EP, Norris F, et al. Zyme, a novel and potentially amyloidogenic enzyme cDNA isolated from Alzheimer's disease brain.
J Biol Chem
1997
;
272
:
25135
-42.
32
Yamashiro K, Tsuruoka N, Kodama S, et al. Molecular cloning of a novel trypsin-like serine protease (neurosin) preferentially expressed in brain.
Biochim Biophys Acta
1997
;
1350
:
11
-4.
33
Yoshida S, Taniguchi M, Hirata A, Shiosaka S. Sequence analysis and expression of human neuropsin cDNA and gene.
Gene
1998
;
213
:
9
-16.
34
Underwood LJ, Tanimoto H, Wang Y, Shigemasa K. Parmley TH, O'Brien TJ. Cloning of tumor-associated differentially expressed gene-14, a novel serine protease overexpressed by ovarian carcinoma.
Cancer Res
1999
;
59
:
4435
-9.
35
Yoshida S, Taniguchi M, Suemoto T, Oka T, He X, Shiosaka S. cDNA cloning and expression of a novel serine protease, TLSP.
Biochim Biophys Acta
1998
;
1399
:
225
-8.
36
Mitsui S, Yamada T, Okui A, Kominami K, Uemura H, Yamaguchi N. A novel isoform of a kallikrein-like protease, TLSP/hippostasin, (PRSS20), is expressed in the human brain and prostate.
Biochem Biophys Res Commun
2000
;
272
:
205
-11.
37
Diamandis EP, Yousef GM, Luo LY, Magklara A, Obiezu CV. The new human kallikrein gene family: implications in carcinogenesis.
Trends Endocrinol Metab
2000
;
11
:
54
-60.
38
Yousef GM, Luo LY, Diamandis EP. Identification of novel human kallikrein-like genes on chromosome 19q13.3-q13.4.
Anticancer Res
1999
;
19
:
2843
-52.
39
Luo L, Herbrick JA, Scherer SW, Beatty B, Squire J, Diamandis EP. Structural characterization and mapping of the normal epithelial cell-specific 1 gene.
Biochem Biophys Res Commun
1998
;
247
:
580
-6.
40
Yousef GM, Diamandis EP. The expanded human kallikrein gene family: locus characterization and molecular cloning of a new member, KLK-L3 (KLK9).
Genomics
2000
;
65
:
184
-94.
41
Yousef GM, Luo LY, Scherer SW, Sotiropoulou G, Diamandis EP. Molecular characterization of Zyme/Protease M/Neurosin (PRSS9), a hormonally regulated kallikrein-like serine protease.
Genomics
1999
;
62
:
251
-9.
42
Yousef GM, Scorilas A, Magklara A, Soosaipillai A, Diamandis EP. The KLK7 (PRSS6) gene, encoding for the stratum corneum chymotryptic enzyme is a new member of the human kallikrein gene family—genomic characterization, mapping, tissue expression and hormonal regulation [In Process Citation].
Gene
2000
;
254
:
119
-28.
43
Yousef GM, Scorilas A, Diamandis EP. Genomic organization, mapping, tissue expression, and hormonal regulation of trypsin-like serine protease (TLSP PRSS20), a new member of the human kallikrein gene family.
Genomics
2000
;
63
:
88
-96.
44
Nelson PS, Gan L, Ferguson C, et al. Molecular cloning and characterization of prostase, an androgen-regulated serine protease with prostate-restricted expression.
Proc Natl Acad Sci USA
1999
;
96
:
3114
-9.
45
Stephenson SA, Verity K, Ashworth LK, Clements JA. Localization of a new prostate-specific antigen-related serine protease gene, KLK4, is evidence for an expanded human kallikrein gene family cluster on chromosome 19q13.3-13.4.
J Biol Chem
1999
;
274
:
23210
-4.
46
Yousef GM, Obiezu CV, Luo LY, Black MH, Diamandis EP. Prostase/KLK-L1 is a new member of the human kallikrein gene family, is expressed in prostate and breast tissues, and is hormonally regulated.
Cancer Res
1999
;
59
:
4252
-6.
47
Brattsand M, Egelrud T. Purification, molecular cloning, and expression of a human stratum corneum trypsin-like serine protease with possible function in desquamation.
J Biol Chem
1999
;
274
:
30033
-40.
48
Yousef GM, Diamandis EP. The new kallikrein-like gene, KLK-L2. Molecular characterization, mapping, tissue expression, and hormonal regulation.
J Biol Chem
1999
;
274
:
37511
-6.
49
Yousef GM, Chang A, Diamandis EP. Identification and characterization of KLK-L4, a new kallikrein-like gene that appears to be down-regulated in breast cancer tissues.
J Biol Chem
2000
;
275
:
11891
-8.
50
Yousef GM, Magklara A, Diamandis EP. KLK12 is a novel serine protease and a new member of the human kallikrein gene family—differential expression in breast cancer.
Genomics
2000
;
69
:
331
-41.
51
Hooper JD, Bui LT, Rae FK, et al. Identification and characterization of klk14, a novel kallikrein serine protease gene located on human chromosome 19q13.4 and expressed in prostate and skeletal muscle.
Genomics
2001
;
73
:
117
-22.
52
Yousef GM, Magklara A, Chang A, Jung K, Katsaros D, Diamandis EP. Cloning of a new member of the human kallikrein gene family, KLK14, which is down-regulated in different malignancies.
Cancer Res
2001
;
61
:
3425
-31.
53
Takayama TK, Carter CA, Deng T. Activation of prostate-specific antigen precursor (pro-PSA) by prostin, a novel human prostatic serine protease identified by degenerate PCR.
Biochemistry
2001
;
40
:
1679
-87.
54
Yousef GM, Scorilas A, Jung K, Ashworth LK, Diamandis EP. Molecular cloning of the human kallikrein 15 gene (KLK15). Up-regulation in prostate cancer.
J Biol Chem
2001
;
276
:
53
-61.
55
Diamandis EP, Yousef GM, Clements J, et al. New nomenclature for the human tissue kallikrein gene family.
Clin Chem
2000
;
46
:
1855
-8.
56
Berg T, Bradshaw RA, Carretero OA, et al. A common nomenclature for members of the tissue (glandular) kallikrein gene families.
Agents Actions Suppl
1992
;
38
:
19
-25.
57
Movat HZ. The plasma kallikrein-kinin system and its interrelationship with other components of blood. In: Erdos EG, editor. Bradykinin, kallidin and kallikrein. Berlin, Germany: Springer-Verlag; 1979. p. 1-89.
58
Asakai R, Davie EW, Chung DW. Organization of the gene for human factor XI.
Biochemistry
1987
;
26
:
7221
-8.
59
Yousef GM, Chang A, Scorilas A, Diamandis EP. Genomic organization of the human kallikrein gene family on chromosome 19q13.3-q13.4.
Biochem Biophys Res Commun
2000
;
276
:
125
-33.
60
Yousef GM, Diamandis M, Jung K, Diamandis EP. Molecular cloning of a novel human acid phosphatase gene (acpt) that is highly expressed in the testis.
Genomics
2001
;
74
:
385
-95.
61
Foussias G, Yousef GM, Diamandis EP. Identification and molecular characterization of a novel member of the siglec family (SIGLEC9). Genomics 
2000
;
67
:
171
-8.
62
Evans BA, Drinkwater CC, Richards RI. Mouse glandular kallikrein genes. Structure and partial sequence analysis of the kallikrein gene locus.
J Biol Chem
1987
;
262
:
8027
-34.
63
Clements J, Hooper J, Dong Y, Harvey T. The expanded human kallikrein (KLK) gene family: genomic organisation, tissue-specific expression and potential functions.
Biol Chem
2001
;
382
:
5
-14.
64
Gomis-Ruth FX, Bayes A, Sotiropoulou G, et al. The structure of human prokallikrein 6 reveals a novel activation mechanism for the kallikrein family.
J Biol Chem
2002
;
277
:
27273
-81.
65
Lu HS, Lin FK, Chao L, Chao J. Human urinary kallikrein. Complete amino acid sequence and sites of glycosylation.
Int J Pept Protein Res
1989
;
33
:
237
-49.
66
Belanger A, van Halbeek H, Graves HC, et al. Molecular mass and carbohydrate structure of prostate specific antigen: studies for establishment of an international PSA standard.
Prostate
1995
;
27
:
187
-97.
67
Mikolajczyk SD, Millar LS, Marker KM, et al. Ala217 is important for the catalytic function and autoactivation of prostate-specific human kallikrein 2.
Eur J Biochem
1997
;
246
:
440
-6.
68
Yousef GM, Kapadia C, Polymeris ME, et al. The human kallikrein protein 5 (hK5) is enzymatically active, glycosylated and forms complexes with two protease inhibitors in ovarian cancer fluids.
Biochim Biophys Acta
2003
;
1628
:
88
-96.
69
Bernett MJ, Blaber SI, Scarisbrick IA, Dhanarajan P, Thompson SM, Blaber M. Crystal structure and biochemical characterization of human kallikrein 6 reveals that a trypsin-like kallikrein is expressed in the central nervous system.
J Biol Chem
2002
;
277
:
24562
-70.
70
Egelrud T. Purification and preliminary characterization of stratum corneum chymotryptic enzyme: a proteinase that may be involved in desquamation.
J Invest Dermatol
1993
;
101
:
200
-4.
71
Kapadia C, Chang A, Sotiropoulou G, et al. Human kallikrein 13: production and purification of recombinant protein and monoclonal and polyclonal antibodies, and development of a sensitive and specific immunofluorometric assay.
Clin Chem
2003
;
49
:
77
-86.
72
Gupta R, Brunak S. Prediction of glycosylation across the human proteome and the correlation to protein function.
Pac Symp Biocomput
2002
;
310
-22.
73
Marshall RD. The nature and metabolism of the carbohydrate-peptide linkages of glycoproteins.
Biochem Soc Symp
1974
;
40
:
17
-26.
74
Matzuk MM, Boime I. The role of the asparagine-linked oligosaccharides of the α subunit in the secretion and assembly of human chorionic gonadotropin.
J Cell Biol
1988
;
106
:
1049
-59.
75
Shakin-Eshleman SH, Remaley AT, Eshleman JR, Wunner WH, Spitalnik SL. N-linked glycosylation of rabies virus glycoprotein. Individual sequons differ in their glycosylation efficiencies and influence on cell surface expression.
J Biol Chem
1992
;
267
:
10690
-8.
76
Scheiffele P, Peranen J, Simons K. N-glycans as apical sorting signals in epithelial cells.
Nature
1995
;
378
:
96
-8.
77
Kadowaki T, Tsukuba T, Bertenshaw GP, Bond JS. N-linked oligosaccharides on the meprin A metalloprotease are important for secretion and enzymatic activity, but not for apical targeting.
J Biol Chem
2000
;
275
:
25577
-84.
78
Hartley BS. Amino-acid sequence of bovine chymotrypsinogen-A.
Nature
1964
;
201
:
1284
-7.
79
Katz BA, Liu B, Barnes M, Springman EB. Crystal structure of recombinant human tissue kallikrein at 2.0 A resolution.
Protein Sci
1998
;
7
:
875
-85.
80
Carvalho AL, Sanz L, Barettino D, Romero A, Calvete JJ, Romao MJ. Crystal structure of a prostate kallikrein isolated from stallion seminal plasma: a homologue of human PSA.
J Mol Biol
2002
;
322
:
325
-37.
81
Kishi T, Kato M, Shimizu T, et al. Crystal structure of neuropsin, a hippocampal protease involved in kindling epileptogenesis.
J Biol Chem
1999
;
274
:
4220
-4.
82
Timm DE. The crystal structure of the mouse glandular kallikrein-13 (prorenin converting enzyme).
Protein Sci
1997
;
6
:
1418
-25.
83
Bode W, Chen Z, Bartels K, Kutzbach C, Schmidt-Kastner G, Bartunik H. Refined 2 A X-ray crystal structure of porcine pancreatic kallikrein A, a specific trypsin-like serine proteinase. Crystallization, structure determination, crystallographic refinement, structure and its comparison with bovine trypsin.
J Mol Biol
1983
;
164
:
237
-82.
84
Matthews BW, Sigler PB, Henderson R, Blow DM. Three-dimensional structure of tosyl-α-chymotrypsin.
Nature
1967
;
214
:
652
-6.
85
Lesk AM, Fordham WD. Conservation and variability in the structures of serine proteinases of the chymotrypsin family.
J Mol Biol
1996
;
258
:
501
-37.
86
Hedstrom L, Szilagyi L, Rutter WJ. Converting trypsin to chymotrypsin: the role of surface loops.
Science
1992
;
255
:
1249
-53.
87
Oka T, Hakoshima T, Itakura M, et al. Role of loop structures of neuropsin in the activity of serine protease and regulated secretion.
J Biol Chem
2002
;
277
:
14724
-30.
88
Varallyay E, Pal G, Patthy A, Szilagyi L, Graf L. Two mutations in rat trypsin confer resistance against autolysis.
Biochem Biophys Res Commun
1998
;
243
:
56
-60.
89
Blaber SI, Scarisbrick IA, Bernett MJ, et al. Enzymatic properties of rat myelencephalon-specific protease.
Biochemistry
2002
;
41
:
1165
-73.
90
Venter JC, Adams MD, Myers EW, et al. The sequence of the human genome.
Science
2001
;
291
:
1304
-51.
91
Lander ES, Linton LM, Birren B, et al. Initial sequencing and analysis of the human genome.
Nature
2001
;
409
:
860
-921.
92
Graveley BR. Alternative splicing: increasing diversity in the proteomic world.
Trends Genet
2001
;
17
:
100
-7.
93
Modrek B, Lee C. A genomic view of alternative splicing.
Nat Genet
2002
;
30
:
13
-9.
94
Johnson JM, Castle J, Garrett-Engele P, et al. Genome-wide survey of human alternative pre-mRNA splicing with exon junction microarrays.
Science
2003
;
302
:
2141
-4.
95
Landry JR, Mager DL, Wilhelm BT. Complex controls: the role of alternative promoters in mammalian genomes.
Trends Genet
2003
;
19
:
640
-8.
96
Beaudoing E, Freier S, Wyatt JR, Claverie JM, Gautheret D. Patterns of variant polyadenylation signal usage in human genes.
Genome Res
2000
;
10
:
1001
-10.
97
Heuze-Vourc'h N, Leblond V, Courty Y. Complex alternative splicing of the hKLK3 gene coding for the tumor marker PSA (prostate-specific-antigen).
Eur J Biochem
2003
;
270
:
706
-14.
98
Burset M, Seledtsov IA, Solovyev VV. SpliceDB: database of canonical and non-canonical mammalian splice sites.
Nucleic Acids Res
2001
;
29
:
255
-9.
99
Thanaraj TA, Clark F. Human GC-AG alternative intron isoforms with weak donor sites show enhanced consensus at acceptor exon positions.
Nucleic Acids Res
2001
;
29
:
2581
-93.
100
Korkmaz KS, Korkmaz CG, Pretlow TG, Saatcioglu F. Distinctly different gene structure of KLK4/KLK-L1/prostase/ARM1 compared with other members of the kallikrein family: intracellular localization, alternative cDNA forms, and regulation by multiple hormones.
DNA Cell Biol
2001
;
20
:
435
-45.
101
Dong Y, Kaushal A, Brattsand M, Nicklin J, Clements JA. Differential splicing of KLK5 and KLK7 in epithelial ovarian cancer produces novel variants with potential as cancer biomarkers.
Clin Cancer Res
2003
;
9
:
1710
-20.
102
Mitsui S, Tsuruoka N, Yamashiro K, Nakazato H, Yamaguchi N. A novel form of human neuropsin, a brain-related serine protease, is generated by alternative splicing and is expressed preferentially in human adult brain.
Eur J Biochem
1999
;
260
:
627
-34.
103
Liu XF, Essand M, Vasmatzis G, Lee B, Pastan I. Identification of three new alternate human kallikrein 2 transcripts: evidence of long transcript and alternative splicing.
Biochem Biophys Res Commun
1999
;
264
:
833
-9.
104
Heuze N, Olayat S, Gutman N, Zani ML, Courty Y. Molecular cloning and expression of an alternative hKLK3 transcript coding for a variant protein of prostate-specific antigen.
Cancer Res
1999
;
59
:
2820
-4.
105
Tanaka T, Isono T, Yoshiki T, Yuasa T, Okada Y. A novel form of prostate-specific antigen transcript produced by alternative splicing.
Cancer Res
2000
;
60
:
56
-9.
106
Pesole G, Mignone F, Gissi C, Grillo G, Licciulli F, Liuni S. Structural and functional features of eukaryotic mRNA untranslated regions.
Gene
2001
;
276
:
73
-81.
107
Mignone F, Gissi C, Liuni S, Pesole G. Untranslated regions of mRNAs.
Genome Biol
2002
;
3(3)
:reviews0004.1-0004.10.
108
Komatsu N, Takata M, Otsuki N, et al. Expression and localization of tissue kallikrein mRNAs in human epidermis and appendages.
J Invest Dermatol
2003
;
121
:
542
-9.
109
Obiezu CV, Soosaipillai A, Jung K, et al. Detection of human kallikrein 4 in healthy and cancerous prostatic tissues by immunofluorometry and immunohistochemistry.
Clin Chem
2002
;
48
:
1232
-40.
110
Yousef GM, Polymeris ME, Grass L, et al. Human kallikrein 5: a potential novel serum biomarker for breast and ovarian cancer.
Cancer Res
2003
;
63
:
3958
-65.
111
Diamandis EP, Yousef GM, Soosaipillai AR, et al. Immunofluorometric assay of human kallikrein 6 (zyme/protease M/neurosin) and preliminary clinical applications.
Clin Biochem
2000
;
33
:
369
-75.
112
Kishi T, Soosaipillai A, Grass L, Little SP, Johnstone EM, Diamandis EP. Development of an immunofluorometric assay and quantification of human kallikrein 7 in tissue extracts and biological fluids.
Clin Chem
2004
;
50(4)
:
709
-16.
113
Kishi T, Grass L, Soosaipillai A, Shimizu-Okabe C, Diamandis EP. Human kallikrein 8: immunoassay development and identification in tissue extracts and biological fluids.
Clin Chem
2003
;
49
:
87
-96.
114
Luo LY, Grass L, Howarth DJ, Thibault P, Ong H, Diamandis EP. Immunofluorometric assay of human kallikrein 10 and its identification in biological fluids and tissues.
Clin Chem
2001
;
47
:
237
-46.
115
Diamandis EP, Okui A, Mitsui S, et al. Human kallikrein 11: a new biomarker of prostate and ovarian carcinoma.
Cancer Res
2002
;
62
:
295
-300.
116
Borgono CA, Grass L, Soosaipillai A, et al. Human kallikrein 14: a new potential biomarker for ovarian and breast cancer.
Cancer Res
2003
;
63
:
9032
-41.
117
Hakalahti L, Vihko P, Henttu P, Autio-Harmainen H, Soini Y, Vihko R. Evaluation of PAP and PSA gene expression in prostatic hyperplasia and prostatic carcinoma using northern-blot analyses, in situ hybridization and immunohistochemical stainings with monoclonal and bispecific antibodies.
Int J Cancer
1993
;
55
:
590
-7.
118
Howarth DJ, Aronson IB, Diamandis EP. Immunohistochemical localization of prostate-specific antigen in benign and malignant breast tissues.
Br J Cancer
1997
;
75
:
1646
-51.
119
Dong Y, Kaushal A, Bui L, et al. Human kallikrein 4 (KLK4) is highly expressed in serous ovarian carcinomas.
Clin Cancer Res
2001
;
7
:
2363
-71.
120
Petraki CD, Karavana VN, Skoufogiannis PT, et al. The spectrum of human kallikrein 6 (zyme/protease M/neurosin) expression in human tissues as assessed by immunohistochemistry.
J Histochem Cytochem
2001
;
49
:
1431
-41.
121
Petraki CD, Karavana VN, Luo LY, Diamandis EP. Human kallikrein 10 expression in normal tissues by immunohistochemistry.
J Histochem Cytochem
2002
;
50
:
1247
-61.
122
Petraki CD, Gregorakis AK, Papanastasiou PA, Karavana VN, Luo LY, Diamandis EP. Immunohistochemical localization of human kallikreins 6, 10 and 13 in benign and malignant prostatic tissues.
Prostate Cancer Prostatic Dis
2003
;
6
:
223
-7.
123
Tanimoto H, Underwood LJ, Shigemasa K, et al. The stratum corneum chymotryptic enzyme that mediates shedding and desquamation of skin cells is highly overexpressed in ovarian tumor cells.
Cancer
1999
;
86
:
2074
-82.
124
Yousef GM, Kyriakopoulou LG, Scorilas A, et al. Quantitative expression of the human kallikrein gene 9 (KLK9) in ovarian cancer: a new independent and favorable prognostic marker.
Cancer Res
2001
;
61
:
7811
-8.
125
Dhar S, Bhargava R, Yunes M, et al. Analysis of normal epithelial cell specific-1 (NES1)/Kallikrein 10 mRNA expression by in situ hybridization, a novel marker for breast cancer.
Clin Cancer Res
2001
;
7
:
3393
-8.
126
Nakamura T, Mitsui S, Okui A, et al. Alternative splicing isoforms of hippostasin (PRSS20/KLK11) in prostate cancer cell lines.
Prostate
2001
;
49
:
72
-8.
127
Petraki CD, Karavana VN, Diamandis EP. Human kallikrein 13 expression in normal tissues: an immunohistochemical study.
J Histochem Cytochem
2003
;
51
:
493
-501.
128
Raidoo DM, Ramsaroop R, Naidoo S, Bhoola KD. Regional distribution of tissue kallikrein in the human brain.
Immunopharmacology
1996
;
32
:
39
-47.
129
Sondell B, Thornell LE, Stigbrand T, Egelrud T. Immunolocalization of stratum corneum chymotryptic enzyme in human skin and oral epithelium with monoclonal antibodies: evidence of a proteinase specifically expressed in keratinizing squamous epithelia.
J Histochem Cytochem
1994
;
42
:
459
-65.
130
Sondell B, Dyberg P, Anneroth GK, Ostman PO, Egelrud T. Association between expression of stratum corneum chymotryptic enzyme and pathological keratinization in human oral mucosa.
Acta Derm Venereol
1996
;
76
:
177
-81.
131
Ekholm IE, Brattsand M, Egelrud T. Stratum corneum tryptic enzyme in normal epidermis: a missing link in the desquamation process?
J Invest Dermatol
2000
;
114
:
56
-63.
132
David A, Mabjeesh N, Azar I, et al. Unusual alternative splicing within the human kallikrein genes KLK2 and KLK3 gives rise to novel prostate-specific proteins.
J Biol Chem
2002
;
277
:
18084
-90.
133
Chang A, Yousef GM, Jung K, Meyts ER, Diamanids EP. Identification and molecular characterization of five novel kallikrein gene 13 (KLK13;KLK-L4) splice variants: differential expression in human testis and testicular cancer.
Anticancer Res
2001
;
21
:
3147
-52.
134
Lillie JW, Green MR. Gene transcription: activator's target in sight.
Nature
1989
;
341
:
279
-80.
135
Trapman J, Cleutjens KB. Androgen-regulated gene expression in prostate cancer.
Semin Cancer Biol
1997
;
8
:
29
-36.
136
Cato AC, Peterziel H. The androgen receptor as mediator of gene expression and signal transduction pathways.
Trends Endocrinol Metab
1998
;
9
:
150
-4.
137
Klein-Hitpass L, Schwerk C, Kahmann S, Vassen L. Targets of activated steroid hormone receptors: basal transcription factors and receptor interacting proteins.
J Mol Med
1998
;
76
:
490
-6.
138
Henderson BE, Feigelson HS. Hormonal carcinogenesis.
Carcinogenesis
2000
;
21
:
427
-33.
139
Riegman PH, Vlietstra RJ, van der Korput HA, Romijn JC, Trapman J. Identification and androgen-regulated expression of two major human glandular kallikrein-1 (hGK-1) mRNA species.
Mol Cell Endocrinol
1991
;
76
:
181
-90.
140
Riegman PH, Vlietstra RJ, van der Korput JA, Brinkmann AO, Trapman J. The promoter of the prostate-specific antigen gene contains a functional androgen responsive element.
Mol Endocrinol
1991
;
5
:
1921
-30.
141
Murtha P, Tindall DJ, Young CY. Androgen induction of a human prostate-specific kallikrein, hKLK2: characterization of an androgen response element in the 5′ promoter region of the gene.
Biochemistry
1993
;
32
:
6459
-64.
142
Cleutjens KB, van der Korput HA, Ehren-van Eekelen CC, et al. A 6-kb promoter fragment mimics in transgenic mice the prostate-specific and androgen-regulated expression of the endogenous prostate-specific antigen gene in humans.
Mol Endocrinol
1997
;
11
:
1256
-65.
143
Young CY, Andrews PE, Tindall DJ. Expression and androgenic regulation of human prostate-specific kallikreins.
J Androl
1995
;
16
:
97
-9.
144
Myers SA, Clements JA. Kallikrein 4 (KLK4), a new member of the human kallikrein gene family is up-regulated by estrogen and progesterone in the human endometrial cancer cell line, KLE.
J Clin Endocrinol & Metab
2001
;
86
:
2323
-6.
145
Luo LY, Grass L, Diamandis EP. The normal epithelial cell-specific 1 (NES1) gene is up-regulated by steroid hormones in the breast carcinoma cell line BT-474.
Anticancer Res
2000
;
20
:
981
-6.
146
Luo LY, Grass L, Diamandis EP. Steroid hormone regulation of the human kallikrein 10 (KLK10) gene in cancer cell lines and functional characterization of the KLK10 gene promoter.
Clin Chim Acta
2003
;
337
:
115
-26.
147
Yousef GM, Fracchioli S, Scorilas A, et al. Steroid hormone regulation and prognostic value of the human kallikrein gene 14 in ovarian cancer.
Am J Clin Pathol
2003
;
119
:
346
-55.
148
Clements J, Mukhtar A, Ehrlich A, Yap B. Glandular kallikrein gene expression in the human uterus.
Braz J Med Biol Res
1994
;
27
:
1855
-63.
149
Aranda A, Pascual A. Nuclear hormone receptors and gene expression.
Physiol Rev
2001
;
81
:
1269
-304.
150
Murray SR, Chao J, Lin FK, Chao L. Kallikrein multigene families and the regulation of their expression.
J Cardiovasc Pharmacol
1990
;
15
:
7
-16.
151
Yu DC, Sakamoto GT, Henderson DR. Identification of the transcriptional regulatory sequences of human kallikrein 2 and their use in the construction of calydon virus 764, an attenuated replication competent adenovirus for prostate cancer therapy.
Cancer Res
1999
;
59
:
1498
-504.
152
Cleutjens KB, van Eekelen CC, van der Korput HA, Brinkmann AO, Trapman J. Two androgen response regions cooperate in steroid hormone regulated activity of the prostate-specific antigen promoter.
J Biol Chem
1996
;
271
:
6379
-88.
153
Cleutjens KB, van der Korput HA, van Eekelen CC, van Rooij HC, Faber PW, Trapman J. An androgen response element in a far upstream enhancer region is essential for high, androgen-regulated activity of the prostate-specific antigen promoter.
Mol Endocrinol
1997
;
11
:
148
-61.
154
Schuur ER, Henderson GA, Kmetec LA, Miller JD, Lamparski HG, Henderson DR. Prostate-specific antigen expression is regulated by an upstream enhancer.
J Biol Chem
1996
;
271
:
7043
-51.
155
Pang S, Dannull J, Kaboo R, et al. Identification of a positive regulatory element responsible for tissue-specific expression of prostate-specific antigen.
Cancer Res
1997
;
57
:
495
-9.
156
Brookes DE, Zandvliet D, Watt F, Russell PJ, Molloy PL. Relative activity and specificity of promoters from prostate-expressed genes.
Prostate
1998
;
35
:
18
-26.
157
Huang W, Shostak Y, Tarr P, Sawyers C, Carey M. Cooperative assembly of androgen receptor into a nucleoprotein complex that regulates the prostate-specific antigen enhancer.
J Biol Chem
1999
;
274
:
25756
-68.
158
Yousef GM, Scorilas A, Magklara A, et al. The androgen-regulated gene human kallikrein 15 (KLK15) is an independent and favourable prognostic marker for breast cancer.
Br J Cancer
2002
;
87
:
1294
-300.
159
McKenna NJ, Xu J, Nawaz Z, Tsai SY, Tsai MJ, O'Malley BW. Nuclear receptor coactivators: multiple enzymes, multiple complexes, multiple functions.
J Steroid Biochem Mol Biol
1999
;
69
:
3
-12.
160
Magklara A, Brown TJ, Diamandis EP. Characterization of androgen receptor and nuclear receptor co-regulator expression in human breast cancer cell lines exhibiting differential regulation of kallikreins 2 and 3.
Int J Cancer
2002
;
100
:
507
-14.
161
Sadar MD. Androgen-independent induction of prostate-specific antigen gene expression via cross-talk between the androgen receptor and protein kinase A signal transduction pathways.
J Biol Chem
1999
;
274
:
7777
-83.
162
Murtha PE, Zhu W, Zhang J, Zhang S, Young CY. Effects of Ca++ mobilization on expression of androgen-regulated genes: interference with androgen receptor-mediated transactivation by AP-I proteins.
Prostate
1997
;
33
:
264
-70.
163
Sun Z, Pan J, Balk SP. Androgen receptor-associated protein complex binds upstream of the androgen-responsive elements in the promoters of human prostate-specific antigen and kallikrein 2 genes.
Nucleic Acids Res
1997
;
25
:
3318
-25.
164
Gaberc-Porekar V, Menart V. Perspectives of immobilized-metal affinity chromatography.
J Biochem Biophys Methods
2001
;
49
:
335
-60.
165
Wang C, Yeung F, Liu PC, et al. Identification of a novel transcription factor, GAGATA-binding protein, involved in androgen-mediated expression of prostate-specific antigen.
J Biol Chem
2003
;
278
:
32423
-30.
166
Cinar B, Yeung F, Konaka H, et al. Identification of a negative regulatory cis-element in the enhancer core region of the Prostate Specific Antigen (PSA) promoter: implications for intersection of androgen receptor and NF-κB signaling in prostate cancer cells.
Biochem J
2004
;
379
:
421
-31.
167
Li B, Goyal J, Dhar S, et al. CpG methylation as a basis for breast tumor-specific loss of NES1/kallikrein 10 expression.
Cancer Res
2001
;
61
:
8014
-21.
168
Roman-Gomez J, Jimenez-Velasco A, Agirre X, et al. The normal epithelial cell-specific 1 (NES1) gene, a candidate tumor suppressor gene on chromosome 19q13.3-4, is downregulated by hypermethylation in acute lymphoblastic leukemia.
Leukemia
2004
;
18
:
362
-5.
169
Li Q, Harju S, Peterson KR. Locus control regions: coming of age at a decade plus.
Trends Genet
1999
;
15
:
403
-8.
170
Li Q, Peterson KR, Fang X, Stamatoyannopoulos G. Locus control regions.
Blood
2002
;
100
:
3077
-86.
171
Smith MS, Lechago J, Wines DR, MacDonald RJ, Hammer RE. Tissue-specific expression of kallikrein family transgenes in mice and rats.
DNA Cell Biol
1992
;
11
:
345
-58.
172
Khan AR, James MN. Molecular mechanisms for the conversion of zymogens to active proteolytic enzymes.
Protein Sci
1998
;
7
:
815
-36.
173
Lovgren J, Rajakoski K, Karp M, Lundwall A, Lilja H. Activation of the zymogen form of prostate-specific antigen by human glandular kallikrein 2.
Biochem Biophys Res Commun
1997
;
238
:
549
-55.
174
Lovgren J, Airas K, Lilja H. Enzymatic action of human glandular kallikrein 2 (hK2). Substrate specificity and regulation by Zn2+ and extracellular protease inhibitors.
Eur J Biochem
1999
;
262
:
781
-9.
175
Christensson A, Laurell CB, Lilja H. Enzymatic activity of prostate-specific antigen and its reactions with extracellular serine proteinase inhibitors.
Eur J Biochem
1990
;
194
:
755
-63.
176
Watt KW, Lee PJ, M'Timkulu T, Chan WP, Loor R. Human prostate-specific antigen: structural and functional similarity with serine proteases.
Proc Natl Acad Sci USA
1986
;
83
:
3166
-70.
177
Sensabaugh GF, Blake ET. Seminal plasma protein p30: simplified purification and evidence for identity with prostate specific antigen.
J Urol
1990
;
144
:
1523
-6.
178
Magklara A, Mellati AA, Wasney GA, et al. Characterization of the enzymatic activity of human kallikrein 6: autoactivation, substrate specificity, and regulation by inhibitors.
Biochem Biophys Res Commun
2003
;
307
:
948
-55.
179
Sotiropoulou G, Rogakos V, Tsetsenis T, et al. Emerging interest in the kallikrein gene family for understanding and diagnosing cancer.
Oncol Res
2003
;
13
:
381
-91.
180
Laskowski M Jr, Qasim MA, Lu SM. Canonical protein inhibitors with serine proteases. In: Kleanthous C, editor. Protein-protein recognition: the frontiers in molecular biology series. Chapter 8. Oxford: Oxford University Press; 2000.
181
Laskowski M, Qasim MA. What can the structures of enzyme-inhibitor complexes tell us about the structures of enzyme substrate complexes?
Biochim Biophys Acta
2000
;
1477
:
324
-37.
182
Silverman GA, Bird PI, Carrell RW, et al. The serpins are an expanding superfamily of structurally similar but functionally diverse proteins. Evolution, mechanism of inhibition, novel functions, and a revised nomenclature.
J Biol Chem
2001
;
276
:
33293
-6.
183
Huntington JA, Read RJ, Carrell RW. Structure of a serpin-protease complex shows inhibition by deformation.
Nature
2000
;
407
:
923
-6.
184
Stenman UH, Leinonen J, Alfthan H, Rannikko S, Tuhkanen K, Alfthan O. A complex between prostate-specific antigen and α1-antichymotrypsin is the major form of prostate-specific antigen in serum of patients with prostatic cancer: assay of the complex improves clinical sensitivity for cancer.
Cancer Res
1991
;
51
:
222
-6.
185
Grauer LS, Finlay JA, Mikolajczyk SD, Pusateri KD, Wolfert RL. Detection of human glandular kallikrein, hK2, as its precursor form and in complex with protease inhibitors in prostate carcinoma serum.
J Androl
1998
;
19
:
407
-11.
186
Becker C, Piironen T, Kiviniemi J, Lilja H, Pettersson K. Sensitive and specific immunodetection of human glandular kallikrein 2 in serum.
Clin Chem
2000
;
46
:
198
-206.
187
Chao J, Chai KX, Chen LM, et al. Tissue kallikrein-binding protein is a serpin. I. Purification, characterization, and distribution in normotensive and spontaneously hypertensive rats.
J Biol Chem
1990
;
265
:
16394
-401.
188
Zhou GX, Chao L, Chao J. Kallistatin: a novel human tissue kallikrein inhibitor. Purification, characterization, and reactive center sequence.
J Biol Chem
1992
;
267
:
25873
-80.
189
Hsieh MC, Cooperman BS. Inhibition of prostate-specific antigen (PSA) by α(1)-antichymotrypsin: salt-dependent activation mediated by a conformational change.
Biochemistry
2002
;
41
:
2990
-7.
190
Sottrup-Jensen L. α-Macroglobulins: structure, shape, and mechanism of proteinase complex formation.
J Biol Chem
1989
;
264
:
11539
-42.
191
Bhoola KD, Figueroa CD, Worthy K. Bioregulation of kinins: kallikreins, kininogens, and kininases.
Pharmacol Rev
1992
;
44
:
1
-80.
192
Clements J. The molecular biology of the kallikreins and their roles in inflammation. In: Farmer S, editor. New York: The Kinin System Academic Press; 1997. p. 71-97.
193
Valdes G, Chacon C, Corthorn J, Figueroa CD, Germain AM. Tissue kallikrein in human placenta in early and late gestation.
Endocrine
2001
;
14
:
197
-204.
194
Valdes G, Germain AM, Corthorn J, Chacon C, Figueroa CD, Muller-Esterl W. Tissue kallikrein and bradykinin B2 receptor in human uterus in luteal phase and in early and late gestation.
Endocrine
2001
;
16
:
207
-15.
195
Schachter M. Kallikreins (kininogenases)—a group of serine proteases with bioregulatory actions.
Pharmacol Rev
1979
;
31
:
1
-17.
196
Bothwell MA, Wilson WH, Shooter EM. The relationship between glandular kallikrein and growth factor-processing proteases of mouse submaxillary gland.
J Biol Chem
1979
;
254
:
7287
-94.
197
Mason AJ, Evans BA, Cox DR, Shine J, Richards RI. Structure of mouse kallikrein gene family suggests a role in specific processing of biologically active peptides.
Nature
1983
;
303
:
300
-7.
198
Jones TH, Figueroa CD, Smith C, Cullen DR, Bhoola KD. Characterization of a tissue kallikrein in human prolactin-secreting adenomas.
J Endocrinol
1990
;
124
:
327
-31.
199
Deperthes D, Marceau F, Frenette G, Lazure C, Tremblay RR, Dube JY. Human kallikrein hK2 has low kininogenase activity while prostate-specific antigen (hK3) has none.
Biochim Biophys Acta
1997
;
1343
:
102
-6.
200
Lovgren J, Tian S, Lundwall A, Karp M, Lilja H. Production and activation of recombinant hK2 with propeptide mutations resulting in high expression levels.
Eur J Biochem
1999
;
266
:
1050
-5.
201
Denmeade SR, Lovgren J, Khan SR, Lilja H, Isaacs JT. Activation of latent protease function of pro-hK2, but not pro-PSA, involves autoprocessing.
Prostate
2001
;
48
:
122
-6.
202
Kumar A, Mikolajczyk SD, Goel AS, Millar LS, Saedi MS. Expression of pro form of prostate-specific antigen by mammalian cells and its conversion to mature, active form by human kallikrein 2.
Cancer Res
1997
;
57
:
3111
-4.
203
Takayama TK, Fujikawa K, Davie EW. Characterization of the precursor of prostate-specific antigen. Activation by trypsin and by human glandular kallikrein.
J Biol Chem
1997
;
272
:
21582
-8.
204
Deperthes D, Frenette G, Brillard-Bourdet M, et al. Potential involvement of kallikrein hK2 in the hydrolysis of the human seminal vesicle proteins after ejaculation.
J Androl
1996
;
17
:
659
-5.
205
Lilja H. A kallikrein-like serine protease in prostatic fluid cleaves the predominant seminal vesicle protein.
J Clin Invest
1985
;
76
:
1899
-903.
206
ole-MoiYoi OK, Pinkus GS, Spragg J, Austen KF. Identification of human glandular kallikrein in the beta cell of the pancreas.
N Engl J Med
1979
;
300
:
1289
-94.
207
Pinkus GS, Maier M, Seldin DC, ole-MoiYoi OK, Austen KF, Spragg J. Immunohistochemical localization of glandular kallikrein in the endocrine and exocrine human pancreas.
J Histochem Cytochem
1983
;
31
:
1279
-88.
208
Petraki CD, Karavana VN, Revelos KI, Luo LY, Diamandis EP. Immunohistochemical localization of human kallikreins 6 and 10 in pancreatic islets.
Histochem J
2002
;
34
:
313
-22.
209
Yoi OO, Seldin DC, Spragg J, Pinkus GS, Austen KF. Sequential cleavage of proinsulin by human pancreatic kallikrein and a human pancreatic kininase.
Proc Natl Acad Sci USA
1979
;
76
:
3612
-6.
210
Lundstrom A, Egelrud T. Cell shedding from human plantar skin in vitro: evidence that two different types of protein structures are degraded by a chymotrypsin-like enzyme.
Arch Dermatol Res
1990
;
282
:
234
-7.
211
Lundstrom A, Egelrud T. Evidence that cell shedding from plantar stratum corneum in vitro involves endogenous proteolysis of the desmosomal protein desmoglein I.
J Invest Dermatol
1990
;
94
:
216
-20.
212
Lundstrom A, Egelrud T. Stratum corneum chymotryptic enzyme: a proteinase which may be generally present in the stratum corneum and with a possible involvement in desquamation.
Acta Derm Venereol
1991
;
71
:
471
-4.
213
Sondell B, Thornell LE, Egelrud T. Evidence that stratum corneum chymotryptic enzyme is transported to the stratum corneum extracellular space via lamellar bodies.
J Invest Dermatol
1995
;
104
:
819
-23.
214
Simon M, Jonca N, Guerrin M, et al. Refined characterization of corneodesmosin proteolysis during terminal differentiation of human epidermis and its relationship to desquamation.
J Biol Chem
2001
;
276
:
20292
-9.
215
Suzuki Y, Koyama J, Moro O, et al. The role of two endogenous proteases of the stratum corneum in degradation of desmoglein-1 and their reduced activity in the skin of ichthyotic patients.
Br J Dermatol
1996
;
134
:
460
-4.
216
Ekholm E, Egelrud T. Stratum corneum chymotryptic enzyme in psoriasis.
Arch Dermatol Res
1999
;
291
:
195
-200.
217
Nylander-Lundqvist E, Back O, Egelrud T. IL-1β activation in human epidermis.
J Immunol
1996
;
157
:
1699
-704.
218
Nylander-Lundqvist E, Egelrud T. Formation of active IL-1β from pro-IL-1β catalyzed by stratum corneum chymotryptic enzyme in vitro.
Acta Derm Venereol
1997
;
77
:
203
-6.
219
Yousef GM, Kishi T, Diamandis EP. Role of kallikrein enzymes in the central nervous system.
Clin Chim Acta
2003
;
329
:
1
-8.
220
Scarisbrick IA, Towner MD, Isackson PJ. Nervous system-specific expression of a novel serine protease: regulation in the adult rat spinal cord by excitotoxic injury.
J Neurosci
1997
;
17
:
8156
-68.
221
Scarisbrick IA, Asakura K, Blaber S, et al. Preferential expression of myelencephalon-specific protease by oligodendrocytes of the adult rat spinal cord white matter.
Glia
2000
;
30
:
219
-30.
222
Scarisbrick IA, Isackson PJ, Ciric B, Windebank AJ, Rodriguez M. MSP, a trypsin-like serine protease, is abundantly expressed in the human nervous system.
J Comp Neurol
2001
;
431
:
347
-61.
223
Scarisbrick IA, Blaber SI, Lucchinetti CF, Genain CP, Blaber M, Rodriguez M. Activity of a newly identified serine protease in CNS demyelination.
Brain
2002
;
125
:
1283
-96.
224
Yamanaka H, He X, Matsumoto K, Shiosaka S, Yoshida S. Protease M/neurosin mRNA is expressed in mature oligodendrocytes.
Brain Res Mol Brain Res
1999
;
71
:
217
-24.
225
Suzuki J, Yoshida S, Chen ZL, et al. Ontogeny of neuropsin mRNA expression in the mouse brain.
Neurosci Res
1995
;
23
:
345
-51.
226
Komai S, Matsuyama T, Matsumoto K, et al. Neuropsin regulates an early phase of schaffer-collateral long-term potentiation in the murine hippocampus.
Eur J Neurosci
2000
;
12
:
1479
-86.
227
Oka T, Akisada M, Okabe A, Sakurai K, Shiosaka S, Kato K. Extracellular serine protease neuropsin (KLK8) modulates neurite outgrowth and fasciculation of mouse hippocampal neurons in culture.
Neurosci Lett
2002
;
321
:
141
-4.
228
Okabe A, Momota Y, Yoshida S, et al. Kindling induces neuropsin mRNA in the mouse brain.
Brain Res
1996
;
728
:
116
-20.
229
Hu JC, Zhang C, Sun X, et al. Characterization of the mouse and human PRSS17 genes, their relationship to other serine proteases, and the expression of PRSS17 in developing mouse incisors.
Gene
2000
;
251
:
1
-8.
230
Hu JC, Sun X, Zhang C, Liu S, Bartlett JD, Simmer JP. Enamelysin and kallikrein-4 mRNA expression in developing mouse molars.
Eur J Oral Sci
2002
;
110
:
307
-15.
231
Simmer JP, Hu JC. Expression, structure, and function of enamel proteinases.
Connect Tissue Res
2002
;
43
:
441
-9.
232
Ryu O, Hu JC, Yamakoshi Y, et al. Porcine kallikrein-4 activation, glycosylation, activity, and expression in prokaryotic and eukaryotic hosts.
Eur J Oral Sci
2002
;
110
:
358
-65.
233
Webber MM, Waghray A, Bello D. Prostate-specific antigen, a serine protease, facilitates human prostate cancer cell invasion.
Clin Cancer Res
1995
;
1
:
1089
-94.
234
Stamenkovic I. Extracellular matrix remodelling: the role of matrix metalloproteinases.
J Pathol
2003
;
200
:
448
-64.
235
Yousef GM, Diamandis EP. Human tissue kallikreins: a new enzymatic cascade pathway?
Biol Chem
2002
;
383
:
1045
-57.
236
Davie EW, Fujikawa K, Kisiel W. The coagulation cascade: initiation, maintenance, and regulation.
Biochemistry
1991
;
30
:
10363
-70.
237
Nesheim M. Thrombin and fibrinolysis.
Chest
2003
;
124
:
33S
-9S.
238
Patthy L. Evolution of the proteases of blood coagulation and fibrinolysis by assembly from modules.
Cell
1985
;
41
:
657
-63.
239
Davie EW, Ichinose A, Leytus SP. Structural features of the proteins participating in blood coagulation and fibrinolysis.
Cold Spring Harbor Symp Quant Biol
1986
;
51 Pt 1
:
509
-14.
240
Yousef GM, Polymeris ME, Yacoub GM, et al. Parallel overexpression of seven kallikrein genes in ovarian cancer.
Cancer Res
2003
;
63
:
2223
-7.
241
Naito K, Fujikawa K. Activation of human blood coagulation factor XI independent of factor XII. Factor XI is activated by thrombin and factor XIa in the presence of negatively charged surfaces.
J Biol Chem
1991
;
266
:
7353
-8.
242
Takayama TK, McMullen BA, Nelson PS, Matsumura M, Fujikawa K. Characterization of hK4 (Prostase), a prostate-specific serine protease: activation of the precursor of prostate specific antigen (pro-PSA) and single-chain urokinase-type plasminogen activator and degradation of prostatic acid phosphatase.
Biochemistry
2001
;
40
:
15341
-8.
243
Frenette G, Tremblay RR, Lazure C, Dube JY. Prostatic kallikrein hK2, but not prostate-specific antigen (hK3), activates single-chain urokinase-type plasminogen activator.
Int J Cancer
1997
;
71
:
897
-9.
244
Choong PF, Nadesapillai AP. Urokinase plasminogen activator system: a multifunctional role in tumor progression and metastasis.
Clin Orthop
2003
;
415 (Suppl)
:
S46
-58.
245
Sternlicht MD, Werb Z. How matrix metalloproteinases regulate cell behavior.
Annu Rev Cell Dev Biol
2001
;
17
:
463
-516.
246
Egeblad M, Werb Z. New functions for the matrix metalloproteinases in cancer progression.
Nat Rev Cancer
2002
;
2
:
161
-74.
247
Desrivieres S, Lu H, Peyri N, Soria C, Legrand Y, Menashi S. Activation of the 92 kDa type IV collagenase by tissue kallikrein.
J Cell Physiol
1993
;
157
:
587
-93.
248
Heidtmann HH, Nettelbeck DM, Mingels A, Jager R, Welker HG, Kontermann RE. Generation of angiostatin-like fragments from plasminogen by prostate-specific antigen.
Br J Cancer
1999
;
81
:
1269
-73.
249
Fortier AH, Holaday JW, Liang H, et al. Recombinant prostate specific antigen inhibits angiogenesis in vitro and in vivo.
Prostate
2003
;
56
:
212
-9.
250
Vu TK, Hung DT, Wheaton VI, Coughlin SR. Molecular cloning of a functional thrombin receptor reveals a novel proteolytic mechanism of receptor activation.
Cell
1991
;
64
:
1057
-68.
251
Macfarlane SR, Seatter MJ, Kanke T, Hunter GD, Plevin R. Proteinase-activated receptors.
Pharmacol Rev
2001
;
53
:
245
-82.
252
Mackie EJ, Pagel CN, Smith R, de Niese MR, Song SJ, Pike RN. Protease-activated receptors: a means of converting extracellular proteolysis into intracellular signals.
IUBMB Life
2002
;
53
:
277
-81.
253
Coughlin SR. Protease-activated receptors in vascular biology.
Thromb Haemost
2001
;
86
:
298
-307.
254
Noorbakhsh F, Vergnolle N, Hollenberg MD, Power C. Proteinase-activated receptors in the nervous system.
Nat Rev Neurosci
2003
;
4
:
981
-90.
255
Ishihara H, Connolly AJ, Zeng D, et al. Protease-activated receptor 3 is a second thrombin receptor in humans.
Nature
1997
;
386
:
502
-6.
256
Xu WF, Andersen H, Whitmore TE, et al. Cloning and characterization of human protease-activated receptor 4.
Proc Natl Acad Sci USA
1998
;
95
:
6642
-6.
257
Nystedt S, Emilsson K, Wahlestedt C, Sundelin J. Molecular cloning of a potential proteinase activated receptor.
Proc Natl Acad Sci USA
1994
;
91
:
9208
-12.
258
Camerer E, Huang W, Coughlin SR. Tissue factor- and factor X-dependent activation of protease-activated receptor 2 by factor VIIa.
Proc Natl Acad Sci USA
2000
;
97
:
5255
-60.
259
Camerer E, Kataoka H, Kahn M, Lease K, Coughlin SR. Genetic evidence that protease-activated receptors mediate factor Xa signaling in endothelial cells.
J Biol Chem
2002
;
277
:
16081
-7.
260
Pendurthi UR, Ngyuen M, Andrade-Gordon P, Petersen LC, Rao LV. Plasmin induces Cyr61 gene expression in fibroblasts via protease-activated receptor-1 and p44/42 mitogen-activated protein kinase-dependent signaling pathway.
Arterioscler Thromb Vasc Biol
2002
;
22
:
1421
-6.
261
Riewald M, Petrovan RJ, Donner A, Mueller BM, Ruf W. Activation of endothelial cell protease activated receptor 1 by the protein C pathway.
Science
2002
;
296
:
1880
-2.
262
Sawada K, Nishibori M, Nakaya N, Wang Z, Saeki K. Purification and characterization of a trypsin-like serine proteinase from rat brain slices that degrades laminin and type IV collagen and stimulates protease-activated receptor-2.
J Neurochem
2000
;
74
:
1731
-8.
263
Obiezu CV, Scorilas A, Katsaros D, et al. Higher human kallikrein gene 4 (klk4) expression indicates poor prognosis of ovarian cancer patients.
Clin Cancer Res
2001
;
7
:
2380
-6.
264
Kim H, Scorilas A, Katsaros D, et al. Human kallikrein gene 5 (KLK5) expression is an indicator of poor prognosis in ovarian cancer.
Br J Cancer
2001
;
84
:
643
-50.
265
Tanimoto H, Underwood LJ, Shigemasa K, Parmley TH, O'Brien TJ. Increased expression of protease M in ovarian tumors.
Tumour Biol
2001
;
22
:
11
-8.
266
Diamandis EP, Yousef GM, Soosaipillai AR, Bunting P. Human kallikrein 6 (zyme/protease M/neurosin): a new serum biomarker of ovarian carcinoma.
Clin Biochem
2000
;
33
:
579
-83.
267
Magklara A, Scorilas A, Katsaros D, et al. The human KLK8 (neuropsin/ovasin) gene: identification of two novel splice variants and its prognostic value in ovarian cancer.
Clin Cancer Res
2001
;
7
:
806
-11.
268
Kishi T, Grass L, Soosaipillai A, et al. Human kallikrein 8, a novel biomarker for ovarian carcinoma.
Cancer Res
2003
;
63
:
2771
-4.
269
Luo LY, Katsaros D, Scorilas A, et al. Prognostic value of human kallikrein 10 expression in epithelial ovarian carcinoma.
Clin Cancer Res
2001
;
7
:
2372
-9.
270
Luo L, Bunting P, Scorilas A, Diamandis EP. Human kallikrein 10: a novel tumor marker for ovarian carcinoma?
Clin Chim Acta
2001
;
306
:
111
-8.
271
Luo LY, Katsaros D, Scorilas A, et al. The serum concentration of human kallikrein 10 represents a novel biomarker for ovarian cancer diagnosis and prognosis.
Cancer Res
2003
;
63
:
807
-11.
272
Yousef GM, Scorilas A, Katsaros D, et al. Prognostic value of the human kallikrein gene 15 expression in ovarian cancer.
J Clin Oncol
2003
;
21
:
3119
-26.
273
Yu H, Giai M, Diamandis EP, et al. Prostate-specific antigen is a new favorable prognostic indicator for women with breast cancer.
Cancer Res
1995
;
55
:
2104
-10.
274
Yu H, Diamandis EP, Levesque M, et al. Prostate specific antigen in breast cancer, benign breast disease and normal breast tissue.
Breast Cancer Res Treat
1996
;
40
:
171
-8.
275
Yu H, Levesque MA, Clark GM, Diamandis EP. Prognostic value of prostate-specific antigen for women with breast cancer: a large United States cohort study.
Clin Cancer Res
1998
;
4
:
1489
-97.
276
Goyal J, Smith KM, Cowan JM, Wazer DE, Lee SW, Band V. The role for NES1 serine protease as a novel tumor suppressor.
Cancer Res
1998
;
58
:
4782
-6.
277
Yousef GM, Yacoub GM, Polymeris ME, Popalis C, Soosaipillai A, Diamandis EP. Kallikrein gene downregulation in breast cancer.
Br J Cancer
2004
;
90
:
167
-72.
278
Magklara A, Scorilas A, Stephan C, et al. Decreased concentrations of prostate-specific antigen and human glandular kallikrein 2 in malignant versus nonmalignant prostatic tissue.
Urology
2000
;
56
:
527
-32.
279
Abrahamsson PA, Lilja H, Falkmer S, Wadstrom LB. Immunohistochemical distribution of the three predominant secretory proteins in the parenchyma of hyperplastic and neoplastic prostate glands.
Prostate
1988
;
12
:
39
-46.
280
Pretlow TG, Pretlow TP, Yang B, et al. Tissue concentrations of prostate-specific antigen in prostatic carcinoma and benign prostatic hyperplasia.
Int J Cancer
1991
;
49
:
645
-9.
281
Yousef GM, Scorilas A, Chang A, et al. Down-regulation of the human kallikrein gene 5 (KLK5) in prostate cancer tissues.
Prostate
2002
;
51
:
126
-32.
282
Luo LY, Diamandis EP. Down-regulation of the normal epithelial cell-specific 1 (NES1) gene is associated with unfavorable outcome of prostate cancer.
Clin Biochem
2000
;
33
:
237
.
283
Yousef GM, Stephan C, Scorilas A, et al. Differential expression of the human kallikrein gene 14 (KLK14) in normal and cancerous prostatic tissues.
Prostate
2003
;
56
:
287
-92.
284
Stephan C, Yousef GM, Scorilas A, et al. Quantitative analysis of kallikrein 15 gene expression in prostate tissue.
J Urol
2003
;
169
:
361
-4.
285
Yousef GM, Obiezu CV, Jung K, Stephan C, Scorilas A, Diamandis EP. Differential expression of Kallikrein gene 5 in cancerous and normal testicular tissues.
Urology
2002
;
60
:
714
-8.
286
Luo LY, Rajpert-De Meyts ER, Jung K, Diamandis EP. Expression of the normal epithelial cell-specific 1 (NES1; KLK10) candidate tumour suppressor gene in normal and malignant testicular tissue.
Br J Cancer
2001
;
85
:
220
-4.
287
Luo LY, Yousef G, Diamandis EP. Human tissue kallikreins and testicular cancer.
APMIS
2003
;
111
:
225
-32.
288
Bhattacharjee A, Richards WG, Staunton J, et al. Classification of human lung carcinomas by mRNA expression profiling reveals distinct adenocarcinoma subclasses.
Proc Natl Acad Sci USA
2001
;
98
:
13790
-5.
289
Iacobuzio-Donahue CA, Ashfaq R, Maitra A, et al. Highly expressed genes in pancreatic ductal adenocarcinomas: a comprehensive characterization and comparison of the transcription profiles obtained from three major technologies.
Cancer Res
2003
;
63
:
8614
-22.
290
Nakamura T, Stephan C, Scorilas A, Yousef GM, Jung K, Diamandis EP. Quantitative analysis of hippostasin/KLK11 gene expression in cancerous and noncancerous prostatic tissues.
Urology
2003
;
61
:
1042
-6.
291
Strathdee G, Brown R. Aberrant DNA methylation in cancer: potential clinical interventions.
Exp Rev Mol Med
2002
;
2002
:
1
-17.
292
Wolf WC, Evans DM, Chao L, Chao J. A synthetic tissue kallikrein inhibitor suppresses cancer cell invasiveness.
Am J Pathol
2001
;
159
:
1797
-805.
293
Cohen P, Peehl DM, Lamson G, Rosenfeld RG. Insulin-like growth factors (IGFs), IGF receptors, and IGF-binding proteins in primary cultures of prostate epithelial cells.
J Clin Endocrinol & Metab
1991
;
73
:
401
-7.
294
Killian CS, Corral DA, Kawinski E, Constantine RI. Mitogenic response of osteoblast cells to prostate-specific antigen suggests an activation of latent TGF-β and a proteolytic modulation of cell adhesion receptors.
Biochem Biophys Res Commun
1993
;
192
:
940
-7.
295
Wu HF, Xu LH, Jenzano JW, et al. Expression of tissue kallikrein in normal and SV40-transfected human endometrial stromal cells.
Pathobiology
1993
;
61
:
123
-7.
296
Chen LM, Richards GP, Chao L, Chao J. Molecular cloning, purification and in situ localization of human colon kallikrein.
Biochem J
1995
;
307 Pt 2
:
481
-6.
297
Hermann A, Buchinger P, Rehbock J. Visualization of tissue kallikrein in human breast carcinoma by two-dimensional western blotting and immunohistochemistry.
Biol Chem Hoppe-Seyler
1995
;
376
:
365
-70.
298
Rehbock J, Buchinger P, Hermann A, Figueroa C. Identification of immunoreactive tissue kallikrein in human ductal breast carcinomas.
J Cancer Res Clin Oncol
1995
;
121
:
64
-8.
299
Koshikawa N, Nakamura T, Tsuchiya N, et al. Purification and identification of a novel and four known serine proteinase inhibitors secreted by human glioblastoma cells.
J Biochem (Tokyo)
1996
;
119
:
334
-9.
300
Mahabeer R, Bhoola KD. Kallikrein and kinin receptor genes.
Pharmacol & Ther
2000
;
88
:
77
-89.
301
Koshikawa N, Yasumitsu H, Umeda M, Miyazaki K. Multiple secretion of matrix serine proteinases by human gastric carcinoma cell lines.
Cancer Res
1992
;
52
:
5046
-53.
302
Dlamini Z, Raidoo D, Bhoola K. Visualisation of tissue kallikrein and kinin receptors in oesophageal carcinoma.
Immunopharmacology
1999
;
43
:
303
-10.
303
Matsumura Y, Kimura M, Yamamoto T, Maeda H. Involvement of the kinin-generating cascade in enhanced vascular permeability in tumor tissue.
Jpn J Cancer Res
1988
;
79
:
1327
-34.
304
Maeda H, Wu J, Okamoto T, Maruo K, Akaike T. Kallikrein-kinin in infection and cancer.
Immunopharmacology
1999
;
43
:
115
-28.
305
Plendl J, Snyman C, Naidoo S, Sawant S, Mahabeer R, Bhoola KD. Expression of tissue kallikrein and kinin receptors in angiogenic microvascular endothelial cells.
Biol Chem
2000
;
381
:
1103
-15.
306
Balbay MD, Juang P, Ilansa N, et al. Stable transfection of prostate cancer cell line PC-3 with prostate specific antigen induces apoptosis both in vivo and in vitro.
Proc Am Assoc Cancer Res
1999
;
40
:
225
-6.
307
Lai LC, Erbas H, Lennard TW, Peaston RT. Prostate-specific antigen in breast cyst fluid: possible role of prostate-specific antigen in hormone-dependent breast cancer.
Int J Cancer
1996
;
66
:
743
-6.
308
Fortier AH, Nelson BJ, Grella DK, Holaday JW. Antiangiogenic activity of prostate-specific antigen.
J Natl Cancer Inst
1999
;
91
:
1635
-40.
309
Duffy MJ. Clinical uses of tumor markers: a critical review.
Crit Rev Clin Lab Sci
2001
;
38
:
225
-62.
310
Diamandis EP. Prostate-specific antigen—its usefulness in clinical medicine.
Trends Endocrinol Metab
1998
;
9
:
310
-6.
311
Diamandis EP, Yousef GM. Human tissue kallikrein gene family: a rich source of novel disease biomarkers.
Exp Rev Mol Diagn
2001
;
1
:
182
-90.
312
Diamandis EP, Yousef GM. Human tissue kallikreins: a family of new cancer biomarkers.
Clin Chem
2002
;
48
:
1198
-205.
313
Yousef GM, Diamandis EP. Expanded human tissue kallikrein family—a novel panel of cancer biomarkers.
Tumour Biol
2002
;
23
:
185
-92.
314
Stenman UH. New ultrasensitive assays facilitate studies on the role of human glandular kallikrein (hK2) as a marker for prostatic disease [editorial; comment].
Clin Chem
1999
;
45
:
753
-4.
315
Magklara A, Scorilas A, Catalona WJ, Diamandis EP. The combination of human glandular kallikrein and free prostate-specific antigen (PSA) enhances discrimination between prostate cancer and benign prostatic hyperplasia in patients with moderately increased total PSA.
Clin Chem
1999
;
45
:
1960
-6.
316
Kwiatkowski MK, Recker F, Piironen T, et al. In prostatism patients the ratio of human glandular kallikrein to free PSA improves the discrimination between prostate cancer and benign hyperplasia within the diagnostic “gray zone” of total PSA 4 to 10 ng/mL.
Urology
1998
;
52
:
360
-5.
317
Haese A, Becker C, Noldus J, et al. Human glandular kallikrein 2: a potential serum marker for predicting the organ confined versus non-organ confined growth of prostate cancer.
J Urol
2000
;
163
:
1491
-7.
318
Diamandis EP, Scorilas A, Fracchioli S, et al. Human kallikrein 6 (hK6): a new potential serum biomarker for diagnosis and prognosis of ovarian carcinoma.
J Clin Oncol
2003
;
21
:
1035
-43.
319
Nakamura T, Scorilas A, Stephan C, Jung K, Soosaipillai AR, Diamandis EP. The usefulness of serum human kallikrein 11 for discriminating between prostate cancer and benign prostatic hyperplasia.
Cancer Res
2003
;
63
:
6543
-6.
320
Stephan C, Jung K, Lein M, Sinha P, Schnorr D, Loening SA. Molecular forms of prostate-specific antigen and human kallikrein 2 as promising tools for early diagnosis of prostate cancer.
Cancer Epidemiol
, Biomarkers & Prev 
2000
;
9
:
1133
-47.
321
Rittenhouse HG, Finlay JA, Mikolajczyk SD, Partin AW. Human Kallikrein 2 (hK2) and prostate-specific antigen (PSA): two closely related, but distinct, kallikreins in the prostate.
Crit Rev Clin Lab Sci
1998
;
35
:
275
-368.
322
Peracaula R, Tabares G, Royle L, et al. Altered glycosylation pattern allows the distinction between prostate-specific antigen (PSA) from normal and tumor origins.
Glycobiology
2003
;
13
:
457
-70.
323
Hoffman BR, Katsaros D, Scorilas A, et al. Immunofluorometric quantitation and histochemical localisation of kallikrein 6 protein in ovarian cancer tissue: a new independent unfavourable prognostic biomarker.
Br J Cancer
2002
;
87
:
763
-71.
324
Kyriakopoulou LG, Yousef GM, Scorilas A, et al. Prognostic value of quantitatively assessed KLK7 expression in ovarian cancer.
Clin Biochem
2003
;
36
:
135
-43.
325
Borgoño CA, Fracchioli S, Yousef GM, et al. Favourable prognostic value of human kallikrein 11 (hK11) in patients with ovarian carcinoma.
Int J Cancer
2003
;
106
:
605
-10.
326
Scorilas A, Borgono CA, Harbeck N, et al. Human kallikrein 13 protein in ovarian cancer cytosols: a new favorable prognostic marker.
J Clin Oncol
2004
;
22
:
678
-85.
327
Yousef GM, Scorilas A, Kyriakopoulou LG, et al. Human kallikrein gene 5 (KLK5) expression by quantitative PCR: an independent indicator of poor prognosis in breast cancer.
Clin Chem
2002
;
48
:
1241
-50.
328
Talieri M, Diamandis EP, Gourgiotis D, Mathioudaki K, Scorilas A. Expression analysis of the human kallikrein 7 (KLK7) in breast tumors: a new potential biomarker for prognosis of breast carcinoma.
Thromb Haemost
2004
;
91
:
180
-6.
329
Yousef GM, Borgono CA, Scorilas A, et al. Quantitative analysis of human kallikrein gene 14 expression in breast tumours indicates association with poor prognosis.
Br J Cancer
2002
;
87
:
1287
-93.
330
Yousef G, Scorilas A, Nakamura T, et al. The prognostic value of the human kallikrein gene 9 (KLK9) in breast cancer.
Breast Cancer Res Treat
2003
;
78
:
149
-58.
331
Chang A, Yousef GM, Scorilas A, et al. Human kallikrein gene 13 (KLK13) expression by quantitative RT-PCR: an independent indicator of favourable prognosis in breast cancer.
Br J Cancer
2002
;
86
:
1457
-64.
332
Foekens JA, Diamandis EP, Yu H, et al. Expression of prostate-specific antigen (PSA) correlates with poor response to tamoxifen therapy in recurrent breast cancer.
Br J Cancer
1999
;
79
:
888
-94.
333
Luo LY, Diamandis EP, Look MP, Soosaipillai AP, Foekens JA. Higher expression of human kallikrein 10 in breast cancer tissue predicts tamoxifen resistance.
Br J Cancer
2002
;
86
:
1790
-6.
334
Stege R, Grande M, Carlstrom K, Tribukait B, Pousette A. Prognostic significance of tissue prostate-specific antigen in endocrine-treated prostate carcinomas.
Clin Cancer Res
2000
;
6
:
160
-5.
335
Naor D, Nedvetzki S, Golan I, Melnik L, Faitelson Y. CD44 in cancer.
Crit Rev Clin Lab Sci
2002
;
39
:
527
-79.
336
Wagner KD, Wagner N, Schedl A. The complex life of WT1.
J Cell Sci
2003
;
116
:
1653
-8.
337
Slawin KM, Shariat SF, Nguyen C, et al. Detection of metastatic prostate cancer using a splice variant-specific reverse transcriptase-polymerase chain reaction assay for human glandular kallikrein.
Cancer Res
2000
;
60
:
7142
-8.
338
De Laurentiis M, De Placido S, Bianco AR, Clark GM, Ravdin PM. A prognostic model that makes quantitative estimates of probability of relapse for breast cancer patients.
Clin Cancer Res
1999
;
5
:
4133
-9.
339
Jerez-Aragones JM, Gomez-Ruiz JA, Ramos-Jimenez G, Munoz-Perez J, Alba-Corlejo, E. A combined neural network and decision trees model for prognosis of breast cancer relapse.
Artif Intell Med
2003
;
27
:
45
-63.
340
Burke HB, Goodman PH, Rosen DB, et al. Artificial neural networks improve the accuracy of cancer survival prediction.
Cancer
1997
;
79
:
857
-62.
341
Zhang Z, Barnhill SD, Zhang H, et al. Combination of multiple serum markers using an artificial neural network to improve specificity in discriminating malignant from benign pelvic masses.
Gynecol Oncol
1999
;
73
:
56
-61.
342
Woolas RP, Conaway MR, Xu F, et al. Combinations of multiple serum markers are superior to individual assays for discriminating malignant from benign pelvic masses.
Gynecol Oncol
1995
;
59
:
111
-16.
343
Slim R, Torremocha F, Moreau T, et al. Loss-of-function polymorphism of the human kallikrein gene with reduced urinary kallikrein activity.
J Am Soc Nephrol
2002
;
13
:
968
-76.
344
Herrala A, Kurkela R, Porvari K, Isomaki R, Henttu P, Vihko, P. Human prostate-specific glandular kallikrein is expressed as an active and an inactive protein.
Clin Chem
1997
;
43
:
279
-84.
345
Bharaj BB, Luo LY, Jung K, Stephan C, Diamandis EP. Identification of single nucleotide polymorphisms in the human kallikrein 10 (KLK10) gene and their association with prostate, breast, testicular, and ovarian cancers.
Prostate
2002
;
51
:
35
-41.
346
Tsuyuki D, Grass L, Diamandis EP. Frequent detection of mutations in the 5′ flanking region of the prostate-specific antigen gene in female breast cancer.
Eur J Cancer
1997
;
33
:
1851
-4.
347
Majumdar S, Diamandis EP. The promoter and the enhancer region of the KLK 3 (prostate specific antigen) gene is frequently mutated in breast tumours and in breast carcinoma cell lines.
Br J Cancer
1999
;
79
:
1594
-602.
348
Nam RK, Zhang WW, Trachtenberg J, et al. Single nucleotide polymorphism of the human kallikrein-2 gene highly correlates with serum human kallikrein-2 levels and in combination enhances prostate cancer detection.
J Clin Oncol
2003
;
21
:
2312
-9.
349
Bharaj B, Scorilas A, Diamandis EP, et al. Breast cancer prognostic significance of a single nucleotide polymorphism in the proximal androgen response element of the prostate specific antigen gene promoter [In Process Citation].
Breast Cancer Res Treat
2000
;
61
:
111
-9.
350
Xue W, Irvine RA, Yu MC, Ross RK, Coetzee GA, Ingles SA. Susceptibility to prostate cancer: interaction between genotypes at the androgen receptor and prostate-specific antigen loci.
Cancer Res
2000
;
60
:
839
-41.
351
Medeiros R, Morais A, Vasconcelos A, et al. Linkage between polymorphisms in the prostate specific antigen ARE1 gene region, prostate cancer risk, and circulating tumor cells.
Prostate
2002
;
53
:
88
-94.
352
Gsur A, Preyer M, Haidinger G, et al. Polymorphic CAG repeats in the androgen receptor gene, prostate-specific antigen polymorphism and prostate cancer risk.
Carcinogenesis
2002
;
23
:
1647
-51.
353
Wang LZ, Sato K, Tsuchiya N, et al. Polymorphisms in prostate-specific antigen (PSA) gene, risk of prostate cancer, and serum PSA levels in Japanese population.
Cancer Lett
2003
;
202
:
53
-9.
354
Yang QF, Sakurai T, Shan L, et al. Novel polymorphisms of prostate-specific antigen (PSA) gene associated with PSA mRNA expression in breast cancer.
J Hum Genet
2000
;
45
:
363
-6.
355
Yang Q, Shan L, Segawa N, et al. Novel polymorphisms in prostate specific antigen gene and its association with prostate cancer.
Anticancer Res
2001
;
21
:
197
-200.
356
Hural JA, Friedman RS, McNabb A, Steen SS, Henderson RA, Kalos M. Identification of naturally processed CD4 T cell epitopes from the prostate-specific antigen kallikrein 4 using peptide-based in vitro stimulation.
J Immunol
2002
;
169
:
557
-65.
357
Day CH, Fanger GR, Retter MW, et al. Characterization of KLK4 expression and detection of KLK4-specific antibody in prostate cancer patient sera.
Oncogene
2002
;
21
:
7114
-20.
358
Huber R, Kukla D, Bode W, et al. Structure of the complex formed by bovine trypsin and bovine pancreatic trypsin inhibitor. II. Crystallographic refinement at 1.9 A resolution.
J Mol Biol
1974
;
89
:
73
-101.
359
Evans BA, Yun ZX, Close JA, et al. Structure and chromosomal localization of the human renal kallikrein gene.
Biochemistry
1988
;
27
:
3124
-9.
360
Lundwall A, Lilja H. Molecular cloning of human prostate specific antigen cDNA.
FEBS Lett
1987
;
214
:
317
-22.
361
Riegman PH, Klaassen P, van der Korput JA, Romijn JC, Trapman J. Molecular cloning and characterization of novel prostate antigen cDNA's.
Biochem Biophys Res Commun
1988
;
155
:
181
-8.
362
Sutherland GR, Baker E, Hyland VJ, et al. Human prostate-specific antigen (APS) is a member of the glandular kallikrein gene family at 19q13.
Cytogenet Cell Genet
1988
;
48
:
205
-7.
363
Sueiras-Diaz J, Jones DM, Ashworth D, Horton J, Evans DM, Szelke M. Cleavage of human kininogen fragments at Met-Lys by human tissue kallikrein.
Braz J Med Biol Res
1994
;
27
:
1935
-42.
364
Chagas JR, Portaro FC, Hirata IY, et al. Determinants of the unusual cleavage specificity of lysyl-bradykinin-releasing kallikreins.
Biochem J
1995
;
306 Pt 1
:
63
-9.
365
Bourgeois L, Brillard-Bourdet M, Deperthes D, et al. Serpin-derived peptide substrates for investigating the substrate specificity of human tissue kallikreins hK1 and hK2.
J Biol Chem
1997
;
272
:
29590
-5.
366
Chen VC, Chao L, Chao J. Roles of the P1, P2, and P3 residues in determining inhibitory specificity of kallistatin toward human tissue kallikrein.
J Biol Chem
2000
;
275
:
38457
-66.
367
Reddigari S, Kaplan AP. Cleavage of human high-molecular weight kininogen by purified kallikreins and upon contact activation of plasma.
Blood
1988
;
71
:
1334
-40.
368
Currie MG, Geller DM, Chao J, Margolius HS, Needleman P. Kallikrein activation of a high molecular weight atrial peptide.
Biochem Biophys Res Commun
1984
;
120
:
461
-6.
369
Hecquet C, Tan F, Marcic BM, Erdos EG. Human bradykinin B(2) receptor is activated by kallikrein and other serine proteases.
Mol Pharmacol
2000
;
58
:
828
-36.
370
Espana F, Fink E, Sanchez-Cuenca J, Gilabert J, Estelles A, Witzgall K. Complexes of tissue kallikrein with protein C inhibitor in human semen and urine.
Eur J Biochem
1995
;
234
:
641
-9.
371
Ecke S, Geiger M, Resch I, et al. Inhibition of tissue kallikrein by protein C inhibitor. Evidence for identity of protein C inhibitor with the kallikrein binding protein.
J Biol Chem
1992
;
267
:
7048
-52.
372
Rahman MM, Worthy K, Elson CJ, Fink E, Dieppe PA, Bhoola KD. Inhibitor regulation of tissue kallikrein activity in the synovial fluid of patients with rheumatoid arthritis.
Br J Rheumatol
1994
;
33
:
215
-23.
373
Hirano K, Okumura Y, Hayakawa S, Adachi T, Sugiura M. Inhibition of human tissue kallikrein by α1-proteinase inhibitor.
Hoppe-Seyler Z Physiol Chem
1984
;
365
:
27
-32.
374
Delaria KA, Muller DK, Marlor CW, et al. Characterization of placental bikunin, a novel human serine protease inhibitor.
J Biol Chem
1997
;
272
:
12209
-14.
375
Cloutier SM, Chagas JR, Mach JP, Gygi CM, Leisinger HJ, Deperthes D. Substrate specificity of human kallikrein 2 (hK2) as determined by phage display technology.
Eur J Biochem
2002
;
269
:
2747
-54.
376
Dube JY, Tremblay RR. Biochemistry and potential roles of prostatic kallikrein hK2.
Mol Urol
1997
;
1
:
279
-85.
377
Heeb MJ, Espana F. α2-Macroglobulin and C1-inactivator are plasma inhibitors of human glandular kallikrein.
Blood Cells Mol Dis
1998
;
24
:
412
-9.
378
Mikolajczyk SD, Millar LS, Marker KM, et al. Identification of a novel complex between human kallikrein 2 and protease inhibitor-6 in prostate cancer tissue.
Cancer Res
1999
;
59
:
3927
-30.
379
Mikolajczyk SD, Millar LS, Kumar A, Saedi MS. Prostatic human kallikrein 2 inactivates and complexes with plasminogen activator inhibitor-1.
Int J Cancer
1999
;
81
:
438
-42.
380
Cao Y, Lundwall A, Gadaleanu V, Lilja H, Bjartell A. Anti-thrombin is expressed in the benign prostatic epithelium and in prostate cancer and is capable of forming complexes with prostate-specific antigen and human glandular kallikrein 2.
Am J Pathol
2002
;
161
:
2053
-63.
381
Mikolajczyk SD, Millar LS, Kumar A, Saedi MS. Human glandular kallikrein, hK2, shows arginine-restricted specificity and forms complexes with plasma protease inhibitors.
Prostate
1998
;
34
:
44
-50.
382
Akiyama K, Nakamura T, Iwanaga S, Hara M. The chymotrypsin-like activity of human prostate-specific antigen, γ-seminoprotein.
FEBS Lett
1987
;
225
:
168
-72.
383
Cohen P, Graves HC, Peehl DM, Kamarei M, Giudice LC, Rosenfeld RG. Prostate-specific antigen (PSA) is an insulin-like growth factor binding protein-3 protease found in seminal plasma.
J Clin Endocrinol & Metab
1992
;
75
:
1046
-53.
384
Sutkowski DM, Goode RL, Baniel J, et al. Growth regulation of prostatic stromal cells by prostate-specific antigen.
J Natl Cancer Inst
1999
;
91
:
1663
-9.
385
Cramer SD, Chen Z, Peehl DM. Prostate specific antigen cleaves parathyroid hormone-related protein in the PTH-like domain: inactivation of PTHrP-stimulated cAMP accumulation in mouse osteoblasts.
J Urol
1996
;
156
:
526
-31.
386
Iwamura M, Hellman J, Cockett AT, Lilja H, Gershagen S. Alteration of the hormonal bioactivity of parathyroid hormone-related protein (PTHrP) as a result of limited proteolysis by prostate-specific antigen.
Urology
1996
;
48
:
317
-25.
387
Christensson A, Laurell CB, Lilja H. Enzymatic activity of prostate-specific antigen and its reactions with extracellular serine proteinase inhibitors.
Eur J Biochem
1990
;
194
:
755
-63.
388
Christensson A, Bjork T, Nilsson O, et al. Serum prostate specific antigen complexed to α1-antichymotrypsin as an indicator of prostate cancer.
J Urol
1993
;
150
:
100
-5.
389
Christensson A, Lilja H. Complex formation between protein C inhibitor and prostate-specific antigen in vitro and in human semen.
Eur J Biochem
1994
;
220
:
45
-53.
390
Zhang WM, Finne P, Leinonen J, Vesalainen S, Nordling S, Stenman UH. Measurement of the complex between prostate-specific antigen and α1-protease inhibitor in serum.
Clin Chem
1999
;
45
:
814
-21.
391
McCormack RT, Rittenhouse HG, Finlay JA, et al. Molecular forms of prostate-specific antigen and the human kallikrein gene family: a new era.
Urology
1995
;
45
:
729
-44.
392
Okui A, Kominami K, Uemura H, Mitsui S, Yamaguchi N. Characterization of a brain-related serine protease, neurosin (human kaillikrein 6), in human cerebrospinal fluid.
NeuroReport
2001
;
12
:
1345
-50.
393
Hutchinson S, Luo LY, Yousef GM, Soosaipillai A, Diamandis EP. Purification of human kallikrein 6 from biological fluids and identification of its complex with α(1)-antichymotrypsin.
Clin Chem
2003
;
49
:
746
-51.
394
He XP, Shiosaka S, Yoshida S. Expression of neuropsin in oligodendrocytes after injury to the CNS.
Neurosci Res
2001
;
39
:
455
-62.
395
Chen LM, Murray SR, Chai KX, Chao L, Chao J. Molecular cloning and characterization of a novel kallikrein transcript in colon and its distribution in human tissues.
Braz J Med Biol Res
1994
;
27
:
1829
-38.
396
Rae F, Bulmer B, Nicol D, Clements J. The human tissue kallikreins (KLKs 1-3) and a novel KLK1 mRNA transcript are expressed in a renal cell carcinoma cDNA library.
Immunopharmacology
1999
;
45
:
83
-8.
397
Heuze-Vourc'h N, Leblond V, Olayat S, Gauthier F, Courty Y. Characterization of PSA-RP2, a protein related to prostate-specific antigen and encoded by alternative hKLK3 transcripts.
Eur J Biochem
2001
;
268
:
4408
-13.
398
Obiezu CV, Diamandis EP. An alternatively spliced variant of KLK4 expressed in prostatic tissue.
Clin Biochem
2000
;
33
:
599
-600.
399
Strausberg RL, Feingold EA, Grouse LH, et al. Generation and initial analysis of more than 15,000 full-length human and mouse cDNA sequences.
Proc Natl Acad Sci USA
2002
;
99
:
16899
-903.
400
Nakamura T, Mitsui S, Okui A, Miki T, Yamaguchi N. Molecular cloning and expression of a variant form of hippostasin/KLK11 in prostate.
Prostate
2003
;
54
:
299
-305.
401
Clark HF, Gurney AL, Abaya E, et al. The secreted protein discovery initiative (SPDI), a large-scale effort to identify novel human secreted and transmembrane proteins: a bioinformatics assessment.
Genome Res
2003
;
13
:
2265
-70.
402
Black MH, Diamandis EP. The diagnostic and prognostic utility of prostate-specific antigen for diseases of the breast.
Breast Cancer Res Treat
2000
;
59
:
1
-14.
403
Partin AW, Catalona WJ, Finlay JA, et al. Use of human glandular kallikrein 2 for the detection of prostate cancer: preliminary analysis.
Urology
1999
;
54
:
839
-45.