Human kallikrein-related peptidases (KLKs) are a subgroup of serine proteases that participate in proteolytic pathways and control protein levels in normal physiology as well as in several pathological conditions. Their complex network of stimulatory and inhibitory interactions may induce inflammatory and immune responses and contribute to the neoplastic phenotype through the regulation of several cellular processes, such as proliferation, survival, migration, and invasion. This family of proteases, which includes one of the most useful cancer biomarkers, kallikrein-related peptidase 3 or PSA, also has a protective effect against cancer promoting apoptosis or counteracting angiogenesis and cell proliferation. Therefore, they represent attractive therapeutic targets and may have important applications in clinical oncology. Despite being intensively studied, many gaps in our knowledge on several molecular aspects of KLK functions still exist. This review aims to summarize recent data on their involvement in different processes related to health and disease, in particular those directly or indirectly linked to the neoplastic process.
1. Introduction
Human kallikrein-related peptidases (KLKs) are a subgroup of serine proteases that have important roles in regulating normal physiological functions, such as immune response, skin desquamation, enamel formation, and semen liquefaction, and the corresponding pathological conditions. There is growing evidence in the literature supporting the view that KLKs are also implicated in tumorigenesis by activating proteolytic processes associated with the neoplastic phenotype. The potential mechanisms involved include the modulation of growth factor bioavailability and activation of hormone and protease-activated receptors (PARs) resulting in proliferative signaling pathways, the degradation of extracellular matrix, cleavage of junction proteins and induction of an epithelial-mesenchymal transition (EMT) phenotype leading to increased tumor cell migration and invasion, and the modulation of interactions between cancer cells and their microenvironment promoting angiogenesis and other protumorigenic processes (reviewed by [1–3]).
The potential of KLKs as cancer markers has been suggested for several members of this protease family [2, 4–6], particularly for kallikrein-related peptidase 3 or prostate-specific antigen (PSA) [7]. PSA is well accepted for assessing recurrence risk in patients with prostate cancer, but its predictive power for diagnosis has been questioned, since several factors other than malignancy may be associated with its high levels in serum, such as preanalytical variables, benign diseases, and drugs [8, 9]. Biomarker panels combining PSA and other promising markers, including members of the KLK family, are expected to improve prostate cancer screening and reduce unnecessary treatments, a strategy that may also be used for detection and monitoring of other malignancies and nonmalignant diseases.
In this paper, we review the current knowledge about the evolution and functions of human kallikrein-related peptidases, their substrates, and their role in health and disease, particularly in the context of cancer.
2. The Human Degradome
Protein synthesis is essential for living, metabolically active cells, but its counterpart, protein degradation, is no less important. Proteolytic mechanisms driven by proteases maintain appropriate protein levels and recognize and degrade the misfolded or mislocalized ones. In addition to acting in nonspecific catabolism, proteases are involved in selective cleavages and activations, modulating protein-protein interactions and contributing to cell signaling both as catalytic units and as multicatalytic complexes. Due to their broad-spectrum actions, proteases play critical roles in regulating normal biological processes, including DNA replication and transcription, cell proliferation, differentiation, and apoptosis. When altered, they may facilitate the development of pathological conditions such as inflammatory and degenerative disorders (reviewed by [10]). The importance of these hydrolytic enzymes is reflected by the number of genes already identified in several mammalian species, with more than 500 in human and primates and even more in rodents [11–14].
The complete set of human proteases—named the human degradome—is distributed in aspartic-, threonine-, cysteine-, serine-, and metalloprotease classes according to the chemical group involved in their catalytic activity [15, 16], and the latter three are the most populated classes [10]. Their substrate cleavage patterns may be specific for a single peptide, as in the case of proteases involved in signaling pathways, or common for a broad range of peptides, which is well exemplified by digestive enzymes [17]. Otherwise, inactive proteases or pseudoproteases bind to their cognate substrate without cleaving them, thus exerting a regulatory function [18].
Detailed information on proteases in prokaryotes and eukaryotes, protease families, pseudogenes, the sequences derived from endogenous retroviruses, 3D structures, substrates, and proteolytic events has been accumulated in different databases such as MEROPS [19] and Degradome [20].
3. The Serine Protease Group
Approximately one-third of proteolytic enzymes are serine proteases, usually endopeptidases. These enzymes use the serine residue present in their active site as a nucleophile to attack the peptide bond of the substrate [21]. In humans, many serine proteases are involved in extra- and intracellular processes mainly related to food digestion, blood coagulation, and immunity (reviewed by [1, 22]). Although these processes are essential for the purposes of catabolism or selective cleavages required for cell signaling, serine protease activity (as well as that of other proteases) is potentially devastating, and several cellular mechanisms were selected to modulate and keep them within limits. For example, they are stored as inactive zymogens or inside granules and can access the substrates only through controlled actions. In addition, serpins, a superfamily of serine protease inhibitors, antagonize their activities in many metabolic pathways, arresting the proteases into an irreversible complex (reviewed by [1]).
Although tightly controlled, several serine proteases have been associated with human diseases. For example, high granzyme levels (granule-secreted enzymes found in cytotoxic T cells and natural killer cells) have been observed in chronic inflammatory diseases such as rheumatoid arthritis [23], asthma [24], diabetes [25], atherosclerosis [26], and chronic obstructive pulmonary [27] and cardiovascular diseases [28]. They have also been implicated in susceptibility to skin tearing and disorganized collagen as observed in chronic wounds and aged/sun-damaged skin (reviewed by [29]). The role of granzymes in these conditions resides in their ability to cleave many substrates, inducing apoptosis through caspase-dependent and caspase-independent pathways [30]. Their potential to create or destroy autoimmune epitopes [31] and be improperly regulated in chronic wounds or released nonspecifically from immune cell into extracellular spaces also contributes to chronic inflammation or extracellular matrix disorganization [27, 32].
Increased levels of neutrophil proteases such as elastase, cathepsin G, and myeloblastin have also been correlated with the severity of cystic fibrosis and chronic obstructive pulmonary disease [33]. Similarly, tryptase and chymase, two serine proteases stored in mast cell granules, take part in the pathophysiology of asthma [34], psoriasis [35], atherothrombosis [36], and fibrotic [37] and inflammatory kidney diseases [38].
With respect to cancer, several serine proteases have been linked to tumor development and progression by activating proteolytic processes that are associated with the neoplastic phenotype (reviewed by [1]). Specifically, a family of serine proteases expressed and secreted in many tissues participates in complex networks of cell signaling pathways that are related to cancer [4–7]. One of the most useful cancer biomarkers in clinical medicine is kallikrein-related peptidase 3 or PSA, which is a member of this family (reviewed by [7]), and there is evidence that other KLKs are also deregulated in cancer and other diseases [4, 39–147] as summarized in Table 1.
Table 1.
Kallikrein-related peptidases. Gene expression pattern, SNPs, and promoter methylation related to cancer and other diseases. CSF = cerebrospinal fluid.
Human kallikreins, initially detected at high levels in pancreas, kallikreas in Greek, include plasma and tissue serine proteases, which are two categories that differ in molecular weight, substrate specificity, and gene structure. The unique plasma kallikrein (PKK) is a glycoprotein encoded by the KLKB1 gene on chromosome region 4q35 and is predominantly synthesized in the liver as an inactive precursor. After activation by the coagulation factor XII, PKK cleaves high molecular weight kininogen to release bradykinin, a mediator of blood coagulation, inflammation, blood pressure, and thrombosis risk [148].
4.1. Kallikrein-Related Peptidases at DNA Level: Genomic Organization and Evolutionary Aspects
The 15 tissue kallikreins or kallikrein-related peptidases (KLKs) are encoded by genes that are tightly clustered in an approximately 300 kb sequence of the 19q13.33–13.41 chromosome region, all containing 5 coding exons with comparable lengths and sequence homology [149, 150]. A pseudogene (KLKP1) has also been assigned to this region [151], as well as multiple repetitive elements such as ALU, Tigger2, MER8, and MSR1 [152]. The large contiguous human KLK gene cluster is limited by the ACPT (testicular acid phosphatase) gene and the Siglec (sialic acid-binding immunoglobulin-like lectin) family of genes at centromeric and telomeric positions, respectively, and other less characterized genes (SNORD88C, C19orf48, MGC45922, and CTU1) (Figure 1).
KLK gene cluster and schematic representation of the human KLK gene and protein structure. (a) KLK gene cluster on the 19q13.33–13.41 chromosome region including the pseudogene KLKP1 and the transcriptional direction from centromere to telomere, except for KLK2 and KLK3, which have the opposite transcriptional direction. The classic KLK genes (KLKs 1–3) are turquoise, KLK4–KLK15 are medium purple, and the Ψ KLK1 processed pseudogene is silver; the arrowheads represent the neighboring genes: ACPT (testicular acid phosphatase) and the Siglec (sialic acid-binding immunoglobulin-like lectin) gene family as well as other less characterized genes (SNORDs, C19orf48, and CTU1). (b) The human KLK gene consists of 5 coding exons (orange boxes represent coding exons; silver boxes represent noncoding exons) and their 4 intervening introns. The positions of the catalytic residues are highly conserved with the histidine (H), aspartic acid (D) 3, and serine (S) codons on coding exons 2, 3, and 5, respectively. Most KLK genes demonstrate alternative splicing, which generates several transcript variants. Alternative 3′ splice sites or skipped exons (shown in green) result in short variants of KLKs 2, 3, 5, 8, and 12 genes. Alternative 5′ splice sites or start sites (shown in blue) also generate short variants of KLKs 2, 3, 6, 7, 8, and 11 genes. Utilization of the alternative exon 6 generates a long transcript encoding a variant of KLK12 gene (shown in blue). (c) KLK proteins are single-chain proteases that are synthesized as preproenzymes and are proteolytically processed to pro-KLKs and secreted after removal of the terminal signal peptide (Pre). The KLK sequence also includes a propeptide (Pro) that maintains the inactive state of the enzyme, as well as a serine protease domain.
KLK gene cluster and schematic representation of the human KLK gene and protein structure. (a) KLK gene cluster on the 19q13.33–13.41 chromosome region including the pseudogene KLKP1 and the transcriptional direction from centromere to telomere, except for KLK2 and KLK3, which have the opposite transcriptional direction. The classic KLK genes (KLKs 1–3) are turquoise, KLK4–KLK15 are medium purple, and the Ψ KLK1 processed pseudogene is silver; the arrowheads represent the neighboring genes: ACPT (testicular acid phosphatase) and the Siglec (sialic acid-binding immunoglobulin-like lectin) gene family as well as other less characterized genes (SNORDs, C19orf48, and CTU1). (b) The human KLK gene consists of 5 coding exons (orange boxes represent coding exons; silver boxes represent noncoding exons) and their 4 intervening introns. The positions of the catalytic residues are highly conserved with the histidine (H), aspartic acid (D) 3, and serine (S) codons on coding exons 2, 3, and 5, respectively. Most KLK genes demonstrate alternative splicing, which generates several transcript variants. Alternative 3′ splice sites or skipped exons (shown in green) result in short variants of KLKs 2, 3, 5, 8, and 12 genes. Alternative 5′ splice sites or start sites (shown in blue) also generate short variants of KLKs 2, 3, 6, 7, 8, and 11 genes. Utilization of the alternative exon 6 generates a long transcript encoding a variant of KLK12 gene (shown in blue). (c) KLK proteins are single-chain proteases that are synthesized as preproenzymes and are proteolytically processed to pro-KLKs and secreted after removal of the terminal signal peptide (Pre). The KLK sequence also includes a propeptide (Pro) that maintains the inactive state of the enzyme, as well as a serine protease domain.
KLK gene cluster and schematic representation of the human KLK gene and protein structure. (a) KLK gene cluster on the 19q13.33–13.41 chromosome region including the pseudogene KLKP1 and the transcriptional direction from centromere to telomere, except for KLK2 and KLK3, which have the opposite transcriptional direction. The classic KLK genes (KLKs 1–3) are turquoise, KLK4–KLK15 are medium purple, and the Ψ KLK1 processed pseudogene is silver; the arrowheads represent the neighboring genes: ACPT (testicular acid phosphatase) and the Siglec (sialic acid-binding immunoglobulin-like lectin) gene family as well as other less characterized genes (SNORDs, C19orf48, and CTU1). (b) The human KLK gene consists of 5 coding exons (orange boxes represent coding exons; silver boxes represent noncoding exons) and their 4 intervening introns. The positions of the catalytic residues are highly conserved with the histidine (H), aspartic acid (D) 3, and serine (S) codons on coding exons 2, 3, and 5, respectively. Most KLK genes demonstrate alternative splicing, which generates several transcript variants. Alternative 3′ splice sites or skipped exons (shown in green) result in short variants of KLKs 2, 3, 5, 8, and 12 genes. Alternative 5′ splice sites or start sites (shown in blue) also generate short variants of KLKs 2, 3, 6, 7, 8, and 11 genes. Utilization of the alternative exon 6 generates a long transcript encoding a variant of KLK12 gene (shown in blue). (c) KLK proteins are single-chain proteases that are synthesized as preproenzymes and are proteolytically processed to pro-KLKs and secreted after removal of the terminal signal peptide (Pre). The KLK sequence also includes a propeptide (Pro) that maintains the inactive state of the enzyme, as well as a serine protease domain.
The colocation and sequence conservation in a wide variety of species make this human tissue serine proteinase family a very interesting target for evolutionary studies [153]. The phylogenetic analysis of KLKs performed by the Maximum Likelihood method [154], using the transcript isoforms of 15 KLK genes, the pseudogene-1 (KLKP1) sequence, and the PRSS1 (trypsin 1) transcript sequence as an external group, reveals five major branches: (a) the classic KLKs (KLKs 1–3), (b) KLKs 4, 5, 7, and 14 and KLKP1, (c) KLKs 9 and 11, (d) KLKs 8, 10, and 15, and (e) KLKs 12 and 13, and a separate branch with KLK6. The tree (Figure 2) is similar in several aspects to other phylogenetic analyses of this cluster [150, 153, 155–157] but also includes the isoforms and reinforces the idea that all KLK genes evolved from a single gene by successive tandem duplications and genomic rearrangements facilitated by repetitive elements.
Phylogenetic relationships within the human tissue KLK gene family in humans. Phylogenetic analysis was performed using the MEGA5 [205] and Maximum Likelihood methods based on the GTR model (General Time Reversible) [154] with Gamma distribution. The bootstrap method was used (with 1000 data set replicates) to investigate node robustness [206]. The phylogenetic tree includes 15 KLK transcripts, the pseudogene-1 (KLKP1) sequence, and the trypsin 1 gene sequence (PRSS1) [155, 156]. The sequences were obtained from the NCBI Reference Sequence (RefSeq) database (http://www.ncbi.nlm.nih.gov/). Numbers indicate the percentage of 1000 bootstrap replicates at each node in the consensus. Bootstrap value ≤95.
The high similarity between KLK2 and KLK3 sequences and the highest support value also suggest that they might have formed by duplication later in evolution. The data grouping KLK4/KLK5 and KLK9/KLK11 also corroborate previous studies [153, 156]. The isolated position of KLK6 in this phylogenetic tree, unlike the findings of other authors, may explain the apparent distance of the remaining family members in respect to normal and pathological functions.
4.2. Kallikrein-Related Peptidases at RNA Level: Transcriptional Regulation Mechanisms
Kallikrein-related peptidase expression is regulated at transcriptional, translational, and posttranslational levels. At the transcriptional level, several response elements (REs) have been identified in the KLK promoters such as an estrogen-related receptor γ (ERRγ) response element [158], a GATA binding motif in KLK1 [159], and functional retinoic acid response elements (RAREs) in KLK10 [160]. Due to the importance of KLK3 expression in prostate cancer, a number of REs have already been described for its promoter, including Sp1/Sp3 [161] and WT1 transcription factor-binding sites [162], a putative p53 RE [163], an XBE (X-factor-binding element that binds specifically to the NF-kappaB p65 subunit) in the AREc (androgen response element enhancer core) [164], and androgen-responsive elements (AREs), the last of which were also present in the KLK2 promoter (reviewed by [127, 165]).
KLK gene expression can also be regulated by epigenetic mechanisms, including histone modifications such as DNA methylation as well as microRNAs (reviewed by [166]), which can affect normal cell physiology and facilitate tumorigenesis if altered. In fact, aberrant promoter methylation leading to KLK10 downregulation has been described in acute lymphoblastic leukemia [167] as well as in breast [168], gastric [91], and prostate cancer [169]. Similarly, abnormal histone acetylation at KLK2 and KLK3 sequences and deregulated expression of miRNAs targeting KLK genes have also been reported in kidney, prostate, and breast cancer cell lines (reviewed by [166]).
In addition to epigenetic events, polymorphisms in regulatory sequences can potentially alter RNA transcription rates and protein levels, as was observed for the homozygous G base substitution (rs266882) in the androgen response element (ARE-1) of the KLK3 promoter [170] and for polymorphic alleles in the 5′-flanking region of the KLK1 gene [171]. KLK gene activity is likewise affected by polymorphisms in the coding region or in the 3′-UTR and downstream sequences of the KLK1, KLK2, KLK3, and KLK7 genes (reviewed by [47]).
According to the NCBI Reference Sequence Database (accessed in November 20, 2014), with the exceptions of KLK1, KLK4, KLK9, KLK13, and KLK14, human KLK genes have multiple isoforms. The alternative transcripts apparently are species specific [155], and a number of them are cancer specific (reviewed by [172]), which supports the idea that they are constantly evolving. The diversity of these isoforms, especially those with no peptidase catalytic motifs, may indicate a type of activity control, for example, by competing for the same substrates or performing different tissue-specific functions [155].
4.3. Kallikrein-Related Peptidases at Protein Level
The KLKs are proteins of 230 amino acids and 28 to 33 kDa, although some small isoforms reach only 3 kDa. Their standard tertiary structure consists of two juxtaposed six-stranded antiparallel β-barrels and two α-helices with the active site between the barrels [173, 174]. They are synthesized as preproenzymes, which are proteolytically processed to pro-KLKs and secreted after removal of the terminal signal peptide. Their ability to release kinins was initially viewed as the definition of a true kallikrein. However, besides plasma kallikrein, only KLK1 has the ability to cleave kininogen (in this case, low molecular weight kininogen) to release kinin. The tissue kallikrein-kinin system can protect against cardiac injury and ischemia/reperfusion-induced cardiomyocyte apoptosis as well as against oxidative stress-induced renal cell apoptosis via stimulation of kinin B2 receptor-Akt [175]. Otherwise, this system appears to be involved in the development of lupus nephritis by increasing local tissue damage triggered by autoimmune inflammation [176] (Figure 3).
Schematic representation of KLK functions related to physiological and pathological conditions. KLKs are involved in several normal processes including blood pressure, coagulation, semen liquefaction, and skin desquamation and can also protect against cardiac injury and ischemia. These proteases may also participate in skin inflammation, neurodegeneration, and autoimmune diseases.
As mentioned above, KLK promoters have several hormone response elements, and their expression can be regulated by steroid hormones [177]. Therefore, KLK levels in different tissues are dependent not only on the presence of specific transcriptional and translational regulators, but also on proteolytic mechanisms, as previously referred to in the degradome section. Shaw and Diamandis [178] detected distinct expression profiles for several kallikrein-related peptidases: KLK1 was highly expressed in the pancreas and salivary gland, KLKs 2, 3 (also observed in seminal plasma), and 11 were highly expressed in the prostate, KLK5 was expressed in the skin, KLK6 was expressed in the brain, KLK9 was expressed in the heart, and KLK12 was expressed in several anatomical sites. KLKs 4, 8, 14, and 15 exhibited a more homogeneous profile or were not detected in various tissues. Komatsu et al. [179] analyzed the skin stratum corneum and identified the presence of many KLKs (KLKs 5–8, 10, 11, 13, and 14). Generally, expression patterns are compatible with their origins—duplicate genes have similar expression patterns in the same tissues, and coexpression patterns are compatible with their physiological functions [153].
5. Kallikrein-Related Peptidases and Their Relationship to Health and Disease
5.1. Normal Physiological Processes and Nonmalignant Diseases
Similar to what has been observed for other proteases, several regulatory mechanisms protect tissues from harmful proteolysis by KLKs. In addition to controlled proenzyme activation and endogenous inhibitors (such as α2-macroglobulin and serpins), there are also inactivating cleavages and allosteric regulation (reviewed by [165]). Regulatory steps may be performed by other proteases including members of the KLK family, which are supported by their coexpression in the same tissue. For example, a KLK cascade including KLK2, KLK14, and probably other KLKs activates pro-KLK3 to generate the mature proteinase that directly cleaves the semenogelins SgI and SgII resulting in seminal clot liquefaction and spermatozoa release [180]. Recently, Yoon et al. [181] observed that MMP-20, which is usually expressed only in dental enamel, processes the prosequence of nine different KLKs and may be a nonspecific activator of the KLK family in pathological conditions.
Another proteolytic cascade has been described for the skin desquamation process in which KLK5 may be autoactivated or activated by KLK14 at neutral pH and then process KLK7, regulating skin desquamation. This cascade may start by KLK6 autoactivation following the cleavage of KLK11, which in turn activates KLK14. Although not completely understood, skin desquamation also depends on other proteases, including cathepsins, aspartic proteases, urokinase, plasmin, and the inflammatory metalloproteinases. Because KLK regulation is critical for proper desquamation, various endogenous inhibitors participate as attenuators of their activities, mainly LEKTI (serine protease inhibitor Kazal-type 5), a protein encoded by the SPINK5 gene. Other factors such as an acidic environment and UV irradiation (and resulting inflammation) may inhibit LEKTI, also contributing to increased KLK expression and enhanced desquamation [64]. The lack of LEKTI expression in Netherton syndrome, a rare genetic skin disease characterized by congenital ichthyosis and severe allergic manifestations, indeed results in increased proteolytic activities of KLK5 and KLK7, which trigger an inflammatory process by activating protease-activated receptor-2 (PAR-2) and stimulating cytokine production [70] (Figure 3).
KLK deregulation is also observed in several other pathological conditions, of which neurodegenerative disorders are good examples (Figure 3). Alzheimer’s disease (AD) and Parkinson’s disease (PD) are the most prevalent human neurodegenerative disorders. Both are caused by the aggregation of proteins: AD is characterized by extracellular deposits of amyloid β (Aβ) and intraneuronal aggregates of tau protein in specific brain regions, and PD is characterized by intracellular neuronal deposits (Lewy bodies and neurites) formed by insoluble α-synuclein [182, 183].
There is convincing evidence from the literature on Alzheimer’s disease that KLK6, the most abundant kallikrein-related peptidase in the central nervous system, cleaves the amyloid precursor protein (APP), a transmembrane glycoprotein from which Aβ derives. The proteolytic activity of KLK6 against APP and substrates in the extracellular matrix and perineuronal net places this peptidase as a potential component of AD pathogenesis. KLK6 expression is reduced in brain tissues, as well as in cerebrospinal fluid of AD patients [42, 184, 185], but the mechanisms behind these findings and their functional consequences are not yet known. Actually, other enzymes (α-, β-, and γ-secretases) cleave APP in different sites and generate several fragments; some of them are aggregation-prone [183]. KLKs may, for example, promote a bias toward synthesis of these toxic fragments by β- and γ-secretases.
Besides KLK6, the kallikrein-related peptidases 7 and 10 show decreased and increased levels, respectively, in cerebrospinal fluid of AD patients [39]. Recently, Shropshire and collaborators observed that KLK7 is able to cleave the core of Aβ in vitro, inhibiting Aβ aggregation and reducing neuronal toxicity [186]. This result may open new opportunities towards treatments for AD.
Several studies on Parkinson’s disease have implicated KLK6 in the degradation of intracellular α-synuclein [187]. Recent data suggested that secreted α-synuclein is also involved in the development of PD by affecting neuronal cell viability [188] and activating inflammatory response [189]. Although still controversial with respect to the intracellular type, KLK6 inefficiency in α-synuclein degradation seems to contribute to PD pathogenesis, probably due to an altered trafficking of KLK6 [187, 190] or to the resistance of certain forms of α-synuclein to KLK6-proteolysis [76, 191].
Multiple sclerosis (MS) is another example of neurodegenerative disorder in which KLK6 levels are altered. In MS patients, KLK6 is abundantly expressed and cleaves myelin proteins, resulting in demyelination and oligodendrogliopathy [192].
As may be noted from AD, PD, and MS data, KLK6 seems to be important for the neuronal homeostasis and survival. However, other kallikrein-related peptidases are probably involved in these processes, as can be deduced from the data on overexpression of KLK1 in epilepsy [193] and on the ability of a set of KLKs (KLK1, KLKs 5–7, and KLK9) to promote neural injury [62].
5.2. Malignant Diseases
As evidenced by the literature, particularly in prostate cancer, KLKs participate in proteolytic pathways that contribute to the neoplastic process (Figure 4). With respect to tumor growth, KLK1 facilitates EGFR and ERK1/2 cascade activation, which is involved in cell proliferation [194]. Similarly, KLK1, KLK2, and KLK3 can regulate tumor growth through IGF-binding protein (IGFBP) degradation, thereby allowing the release of the insulin-like growth factors (IGFs) and proliferative signals. However, a negative regulatory role for KLK3 in cancer has also been suggested because this protease can activate latent transforming growth factor-β (TGFβ), a known suppressor of growth and promoter of apoptosis [2].
Kallikrein-related peptidases and cancer. KLKs participate in proteolytic pathways that contribute to the neoplastic process by facilitating cell proliferation via growth factors and modulating cell survival through mTOR signaling. They can also regulate angiogenesis, cell migration and invasion by angiogenic vascular endothelial growth factor (VEGF) secretion, metalloproteinase activation, extracellular matrix (ECM) degradation, and epithelial-mesenchymal transition (EMT) induction. However, KLKs also have a protective effect against cancer, promoting apoptosis or inhibiting angiogenesis and cell proliferation.
Recent data have demonstrated that kallikrein-related peptidase 4 and its substrate, promyelocytic leukemia zinc finger protein (PLZF), modulate androgen receptor (AR) and mTOR signaling in prostate cells to regulate cell survival. In fact, KLK4 negatively regulates PLZF, thus preventing its binding and inhibition by AR, which keeps mTORC1 signaling active and ensures cell survival [195].
During neoplastic progression, different KLKs can regulate new vessel formation, which are essential to provide oxygen and nutrients to proliferating cancerous cells. KLKs 1 and 4 stimulate angiogenesis by cleaving kininogen to kinin or activating prometalloproteinases 2 and 9 to their active forms, thereby potentiating extracellular matrix hydrolysis and enabling endothelial cell migration and neovascularization [196–198]. Other kallikrein-related peptidases (KLKs 2 and 4) can stimulate the urokinase plasminogen activator (uPA)/uPA receptor system, which also leads to metalloproteinase activation and extracellular matrix degradation [2]. KLK12 may then promote angiogenesis by the conversion of the membrane-bound platelet-derived growth factor B (PDGF-B) precursor into a soluble form that modulates secretion of the angiogenic vascular endothelial growth factor A (VEGF-A) [199]. Some KLKs, such as KLKs 3, 6, and 13, have the opposite action by blocking VEGF and/or fibroblast growth factor 2 (FGF2) or generating angiostatin-like fragments from plasminogen, which are potent inhibitors of angiogenesis in vitro [2].
Tumors have an increased acidic microenvironment resulting from accelerated glycolysis and lactate accumulation and thus low pH in the extracellular space [200]. Because an acidic environment may block the kallikrein inhibitor LEKTI, contributing to increased KLK expression and loss of cellular adhesion in skin desquamation, it is reasonable to consider a similar mechanism during neoplastic dissemination [201]. In fact, the metastatic process is associated with a transition from tightly connected cells to cells with increased motility, namely, the epithelial-mesenchymal transition, where KLKs play important roles. For example, KLKs activate latent TGFβ, which induces EMT, and are associated with the loss of E-cadherin in tumor cells and thus with decreased cell-cell adhesion [202]. They also trigger extracellular matrix degradation via prometalloprotease activation and hence promote tissue invasion [2].
These examples demonstrate how important kallikrein-related peptidases are in tumor development and progression. The biological processes in which they participate are related to diseases other than cancer but are directly connected with cancer pathways, including cell proliferation, adhesion, inflammation, and apoptosis.
6. Therapeutic Relevance of KLKs
As discussed in previous sections, KLKs have been associated with different pathologic processes, from skin diseases to neurodegenerative disorders and cancer. The progress in our knowledge on all members of this protein family, functions, 3D structures, substrates, and physiological roles, has provided opportunities to develop new therapeutic approaches for different disorders.
KLKs are targeted by several types of inhibitors, including small-molecule inhibitors, antibody-, protein-, and peptide-based inhibitors, KLK-activated prodrugs, interfering RNAs, and immunotherapeutic vaccines (reviewed by [3]). PROSTVAC, for example, is a prostate cancer vaccine consisting of a KLK3 recombinant vector that contains transgenes for three T-cell costimulatory molecules (TRICOM). This vaccine has demonstrated success in inhibiting, with few side effects, cell proliferation and tumor growth and in improving overall survival [203].
Prodrugs activated by KLKs are another strategy that has been investigated. For instance, KLK3-activated peptides have the ability to target the prostate since most KLK3 is expressed in the gland whereas circulating KLK3 is normally inactivated in plasma by endogenous inhibitors [204]. This drug has overcome the challenge of specificity, although similar successful results are not always achieved. The reasons for that include the fact that the active sites of members of KLK family are conserved, which hampers drug design. The resolution of 3D structure of KLKs should help in this regard. However, KLKs also have overlapping and even opposing actions, which certainly depend on the physiologic, tissue, and disease context [203].
7. Conclusions
The KLK network is impressive. Its intricate signaling pathways and protein interactions strongly show that this group of proteases contributes to normal and pathological metabolisms. However, despite being intensively studied, there are many gaps in our knowledge on the molecular aspects of the KLK family. For example, there is no doubt that KLK expression deregulation participates in the development of neurodegenerative disorders. But what exactly is its role? Would it be a primary and direct one, promoting erroneously protein degradation, which results in pathogenic fragments? Or would it be one that implies cooperating with specific secretases and other enzymes to generate toxic deposits?
In cancer, it is not clear whether KLKs alterations are driver mutations or deleterious passenger mutations. The fact that similar sets of KLKs are associated with different tumor types and facilitate proliferation, migration, and other cancer hallmarks aligns with driver mutations. Differently, antiproliferative effects of KLKs and similar regulatory factors for different members of this family may argue in favor of random passenger mutations. However, both statements are not mutually exclusive and may occur simultaneously or sequentially. In fact, the idea of sequential occurrence is interesting: considering the complexity of human proteolytic system, it is reasonable to assume that the expression of specific KLKs may counteract the under- or overexpression of other KLKs or enzymes or even that those KLKs are activated, one after the other, to neutralize the expression of a driver mutation, but without success. The analysis of KLK panels in large sets of samples from diverse stages of the disease, including premalignant phases, will probably help to reveal how the expression profile evolves during the course of the disease.
Many questions are still unanswered and the scenario is therefore incomplete. Many more data are necessary to improve our understanding on the function, substrates, and role of KLKs in health and disease in order to distinguish in each case whether they are heroes, villains, or supporting actors.
Conflict of Interests
The authors declare that they have no conflict of interests.
Acknowledgments
The authors thank the Fundação de Amparo à Pesquisa do Estado de São Paulo (FAPESP) (Grant no. FAPESP 10/51168-0); Coordenação de Aperfeiçoamento de Pessoal de Nível Superior (CAPES); and Conselho Nacional de Pesquisas (CNPq) (Grant no. CNPq 306216/2010-8) for financial support and fellowships. They are also grateful to Mauro Golin for artwork preparation and to GENCAPO (Head and Neck Genome Project, http://www.gencapo.famerp.br/) team for the valuable discussions that motivated the present study.
1Heutinck K. M.,
ten Berge I. J. M.,
Hack C. E.,
Hamann J., and
Rowshani A. T., Serine proteases of the human immune system in health and disease, Molecular Immunology. (2010) 47, no. 11-12, 1943–1955, https://doi.org/10.1016/j.molimm.2010.04.020, 2-s2.0-77953609646.
2Borgoño C. A. and
Diamandis E. P., The emerging roles of human tissue kallikreins in cancer, Nature Reviews Cancer. (2004) 4, no. 11, 876–890, https://doi.org/10.1038/nrc1474, 2-s2.0-8144222388.
3Kryza T.,
Silva M. L.,
Loessner D.,
Heuzé-Vourc′h N., and
Clements J. A., The kallikrein-related peptidase family: dysregulation and functions during cancer progression, Biochimie. (2015) https://doi.org/10.1016/j.biochi.2015.09.002.
4Schmitt M.,
Magdolen V.,
Yang F.,
Kiechle M.,
Bayani J.,
Yousef G. M.,
Scorilas A.,
Diamandis E. P., and
Dorn J., Emerging clinical importance of the cancer biomarkers kallikrein-related peptidases (KLK) in female and male reproductive organ malignancies, Radiology and Oncology. (2013) 47, no. 4, 319–329, https://doi.org/10.2478/raon-2013-0053, 2-s2.0-84893806399.
5Dong Y.,
Loessner D.,
Irving-Rodgers H.,
Obermair A.,
Nicklin J. L., and
Clements J. A., Metastasis of ovarian cancer is mediated by kallikrein related peptidases, Clinical and Experimental Metastasis. (2014) 31, no. 1, 135–147, https://doi.org/10.1007/s10585-013-9615-4, 2-s2.0-84895021159.
6Yousef G. M.,
Yacoub G. M.,
Polymeris M.-E.,
Popalis C.,
Soosaipillai A., and
Diamandis E. P., Kallikrein gene downregulation in breast cancer, British Journal of Cancer. (2004) 90, no. 1, 167–172, https://doi.org/10.1038/sj.bjc.6601451, 2-s2.0-0842265103.
7Sotiropoulou G.,
Pampalakis G., and
Diamandis E. P., Functional roles of human Kallikrein-related peptidases, Journal of Biological Chemistry. (2009) 284, no. 48, 32989–32994, https://doi.org/10.1074/jbc.r109.027946, 2-s2.0-70450231593.
8Duffy M. J., PSA in screening for prostate cancer: more good than harm or more harm than good?, Advances in Clinical Chemistry. (2014) 66, 1–23, https://doi.org/10.1016/b978-0-12-801401-1.00001-3, 2-s2.0-84906080518.
9Scorilas A. and
Mavridis K., Predictions for the future of kallikrein-related peptidases in molecular diagnostics, Expert Review of Molecular Diagnostics. (2014) 14, no. 6, 713–722, https://doi.org/10.1586/14737159.2014.928207, 2-s2.0-84903159211.
10López-Otín C. and
Bond J. S., Proteases: multifunctional enzymes in life and disease, The Journal of Biological Chemistry. (2008) 283, no. 45, 30433–30437, https://doi.org/10.1074/jbc.r800035200, 2-s2.0-57649155302.
11Puente X. S.,
Gutiérrez-Fernández A.,
Ordóñez G. R.,
Hillier L. W., and
López-Otín C., Comparative genomic analysis of human and chimpanzee proteases, Genomics. (2005) 86, no. 6, 638–647, https://doi.org/10.1016/j.ygeno.2005.07.009, 2-s2.0-28644432882.
12Puente X. S.,
Sánchez L. M.,
Overall C. M., and
López-Otín C., Human and mouse proteases: a comparative genomic approach, Nature Reviews Genetics. (2003) 4, no. 7, 544–558, https://doi.org/10.1038/nrg1111, 2-s2.0-0037803570.
13Puente X. S. and
López-Otín C., A genomic analysis of rat proteases and protease inhibitors, Genome Research. (2004) 14, no. 4, 609–622, https://doi.org/10.1101/gr.1946304, 2-s2.0-0347445722.
14Quesada V.,
Velasco G.,
Puente X. S.,
Warren W. C., and
López-Otín C., Comparative genomic analysis of the zebra finch degradome provides new insights into evolution of proteases in birds and mammals, BMC Genomics. (2010) 11, article 220, https://doi.org/10.1186/1471-2164-11-220, 2-s2.0-77950482175.
15Fujinaga M.,
Cherney M. M.,
Oyama H.,
Oda K., and
James M. N. G., The molecular structure and catalytic mechanism of a novel carboxyl peptidase from Scytalidium lignicolum, Proceedings of the National Academy of Sciences of the United States of America. (2004) 101, no. 10, 3364–3369, https://doi.org/10.1073/pnas.0400246101, 2-s2.0-1542723495.
16Rawlings N. D.,
Tolle D. P., and
Barrett A. J., MEROPS: the peptidase database, Nucleic Acids Research. (2004) 32, D160–D164, https://doi.org/10.1093/nar/gkh071, 2-s2.0-0345864015.
17Fuchs J. E.,
von Grafenstein S.,
Huber R. G.,
Kramer C., and
Liedl K. R., Substrate-driven mapping of the degradome by comparison of sequence logos, PLoS Computational Biology. (2013) 9, no. 11, e1003353, https://doi.org/10.1371/journal.pcbi.1003353, 2-s2.0-84888274187.
18Bergbold N. and
Lemberg M. K., Emerging role of rhomboid family proteins in mammalian biology and disease, Biochimica et Biophysica Acta. (2013) 1828, no. 12, 2840–2848, https://doi.org/10.1016/j.bbamem.2013.03.025, 2-s2.0-84885179066.
19Rawlings N. D.,
Barrett A. J., and
Bateman A., MEROPS: the database of proteolytic enzymes, their substrates and inhibitors, Nucleic Acids Research. (2012) 40, no. 1, D343–D350, https://doi.org/10.1093/nar/gkr987, 2-s2.0-84859426282.
20Quesada V.,
Ordóñez G. R.,
Sánchez L. M.,
Puente X. S., and
López-Otín C., The degradome database: mammalian proteases and diseases of proteolysis, Nucleic Acids Research. (2009) 37, no. 1, D239–D243, https://doi.org/10.1093/nar/gkn570, 2-s2.0-58149178573.
21Page M. J. and
Di Cera E., Serine peptidases: classification, structure and function, Cellular and Molecular Life Sciences. (2008) 65, no. 7-8, 1220–1236, https://doi.org/10.1007/s00018-008-7565-9, 2-s2.0-42249109450.
23Saito S.,
Murakoshi K.,
Kotake S.,
Kamatani N., and
Tomatsu T., Granzyme B induces apoptosis of chondrocytes with natural killer cell-like cytotoxicity in rheumatoid arthritis, Journal of Rheumatology. (2008) 35, no. 10, 1932–1943, 2-s2.0-54949117926.
24Tschopp C. M.,
Spiegl N.,
Didichenko S.,
Lutmann W.,
Julius P.,
Virchow J. C.,
Hack C. E., and
Dahinden C. A., Granzyme B, a novel mediator of allergic inflammation: its induction and release in blood basophils and human asthma, Blood. (2006) 108, no. 7, 2290–2299, https://doi.org/10.1182/blood-2006-03-010348, 2-s2.0-33749315904.
25Thomas H. E.,
Trapani J. A., and
Kay T. W. H., The role of perforin and granzymes in diabetes, Cell Death and Differentiation. (2010) 17, no. 4, 577–585, https://doi.org/10.1038/cdd.2009.165, 2-s2.0-77949510984.
26Chamberlain C. M. and
Granville D. J., The role of Granzyme B in atheromatous diseases, Canadian Journal of Physiology and Pharmacology. (2007) 85, no. 1, 89–95, https://doi.org/10.1139/y06-090, 2-s2.0-34250885278.
27Ngan D. A.,
Vickerman S. V.,
Granville D. J.,
Man S. F. P., and
Sin D. D., The possible role of granzyme B in the pathogenesis of chronic obstructive pulmonary disease, Therapeutic Advances in Respiratory Disease. (2009) 3, no. 3, 113–129, https://doi.org/10.1177/1753465809341965, 2-s2.0-68549083691.
28Saito Y.,
Kondo H., and
Hojo Y., Granzyme B as a novel factor involved in cardiovascular diseases, Journal of Cardiology. (2011) 57, no. 2, 141–147, https://doi.org/10.1016/j.jjcc.2010.10.001, 2-s2.0-79951944773.
29Hiebert P. R. and
Granville D. J., Granzyme B in injury, inflammation, and repair, Trends in Molecular Medicine. (2012) 18, no. 12, 732–741, https://doi.org/10.1016/j.molmed.2012.09.009, 2-s2.0-84870255907.
30Boivin W. A.,
Cooper D. M.,
Hiebert P. R., and
Granville D. J., Intracellular versus extracellular granzyme B in immunity and disease: challenging the dogma, Laboratory Investigation. (2009) 89, no. 11, 1195–1220, https://doi.org/10.1038/labinvest.2009.91, 2-s2.0-70350627504.
31Darrah E. and
Rosen A., Granzyme B cleavage of autoantigens in autoimmunity, Cell Death and Differentiation. (2010) 17, no. 4, 624–632, https://doi.org/10.1038/cdd.2009.197, 2-s2.0-77949529810.
32Hendel A.,
Hiebert P. R.,
Boivin W. A.,
Williams S. J., and
Granville D. J., Granzymes in age-related cardiovascular and pulmonary diseases, Cell Death and Differentiation. (2010) 17, no. 4, 596–606, https://doi.org/10.1038/cdd.2010.5, 2-s2.0-77949530057.
33Roghanian A. and
Sallenave J.-M., Neutrophil elastase (NE) and NE inhibitors: canonical and noncanonical functions in lung chronic inflammatory diseases (cystic fibrosis and chronic obstructive pulmonary disease), Journal of Aerosol Medicine and Pulmonary Drug Delivery. (2008) 21, no. 1, 125–144, https://doi.org/10.1089/jamp.2007.0653, 2-s2.0-45949091910.
34Fajt M. L. and
Wenzel S. E., Mast cells, their subtypes, and relation to asthma phenotypes, Annals of the American Thoracic Society. (2013) 10, supplement, S158–S164, https://doi.org/10.1513/annalsats.201303-064aw, 2-s2.0-84892525338.
35Namazi M. R., Possible molecular mechanisms to account for the involvement of tryptase in the pathogenesis of psoriasis, Autoimmunity. (2005) 38, no. 6, 449–452, https://doi.org/10.1080/08916930500246289, 2-s2.0-27944474841.
36Kovanen P. T., Mast cells and degradation of pericellular and extracellular matrices: potential contributions to erosion, rupture and intraplaque haemorrhage of atherosclerotic plaques, Biochemical Society Transactions. (2007) 35, no. 5, 857–861, https://doi.org/10.1042/bst0350857, 2-s2.0-36749022372.
39Diamandis E. P.,
Scorilas A.,
Kishi T.,
Blennow K.,
Luo L.-Y.,
Soosaipillai A.,
Rademaker A. W., and
Sjogren M., Altered kallikrein 7 and 10 concentrations in cerebrospinal fluid of patients with Alzheimer′s disease and frontotemporal dementia, Clinical Biochemistry. (2004) 37, no. 3, 230–237, https://doi.org/10.1016/j.clinbiochem.2003.11.012, 2-s2.0-1242269946.
40Diamandis E. P.,
Yousef G. M.,
Petraki C., and
Soosaipillai A. R., Human kallikrein 6 as a biomarker of Alzheimer′s disease, Clinical Biochemistry. (2000) 33, no. 8, 663–667, https://doi.org/10.1016/S0009-9120(00)00185-5, 2-s2.0-0034483227.
41Menendez-Gonzalez M.,
Castro-Santos P.,
Suarez A.,
Calatayud M. T.,
Perez-Pinera P.,
Martinez M.,
Ribacoba R., and
Gutierrez C., Value of measuring plasmatic levels of neurosin in the diagnosis of Alzheimer′s disease, Journal of Alzheimer′s Disease. (2008) 14, no. 1, 59–67, 2-s2.0-47849088938.
42Ashby E. L.,
Kehoe P. G., and
Love S., Kallikrein-related peptidase 6 in Alzheimer′s disease and vascular dementia, Brain Research. (2010) 1363, 1–10, https://doi.org/10.1016/j.brainres.2010.09.017, 2-s2.0-78149468228.
43Wang S.-K.,
Hu Y.,
Simmer J. P.,
Seymen F.,
Estrella N. M. R. P.,
Pal S.,
Reid B. M.,
Yildirim M.,
Bayram M.,
Bartlett J. D., and
Hu J. C.-C., Novel KLK4 and MMP20 mutations discovered by whole-exome sequencing, Journal of Dental Research. (2013) 92, no. 3, 266–271, https://doi.org/10.1177/0022034513475626, 2-s2.0-84874097203.
44Chan H.-C.,
Estrella N. M. R. P.,
Milkovich R. N.,
Kim J.-W.,
Simmer J. P., and
Hu J. C.-C., Target gene analyses of 39 amelogenesis imperfecta kindreds, European Journal of Oral Sciences. (2011) 119, no. supplement 1, 311–323, https://doi.org/10.1111/j.1600-0722.2011.00857.x, 2-s2.0-84862909523.
45Lee S.-K.,
Seymen F.,
Lee K.-E.,
Kang H.-Y.,
Yildirim M.,
Bahar Tuna E.,
Gencay K.,
Hwang Y.-H.,
Nam K. H.,
De La Garza R. J.,
Hu J. C.-C.,
Simmer J. P., and
Kim J.-W., Novel WDR72 mutation and cytoplasmic localization, Journal of Dental Research. (2010) 89, no. 12, 1378–1382, https://doi.org/10.1177/0022034510382117, 2-s2.0-78650085380.
46Martínez-Morillo E.,
Diamandis A.,
Romaschin A. D., and
Diamandis E. P., Kallikrein 6 as a serum prognostic marker in patients with aneurysmal subarachnoid hemorrhage, PLoS ONE. (2012) 7, no. 9, e45676, https://doi.org/10.1371/journal.pone.0045676, 2-s2.0-84866661677.
47Batra J.,
O′Mara T.,
Patnala R.,
Lose F., and
Clements J. A., Genetic polymorphisms in the human tissue kallikrein (KLK) locus and their implication in various malignant and non-malignant diseases, Biological Chemistry. (2012) 393, no. 12, 1365–1390, https://doi.org/10.1515/hsz-2012-0211, 2-s2.0-84869475828.
48Myers R. A.,
Himes B. E.,
Gignoux C. R.,
Yang J. J.,
Gauderman W. J.,
Rebordosa C.,
Xie J.,
Torgerson D. G.,
Levin A. M.,
Baurley J.,
Graves P. E.,
Mathias R. A.,
Romieu I.,
Roth L. A.,
Conti D.,
Avila L.,
Eng C.,
Vora H.,
LeNoir M. A.,
Soto-Quiros M.,
Liu J.,
Celedón J. C.,
Farber H. J.,
Kumar R.,
Avila P. C.,
Meade K.,
Serebrisky D.,
Thyne S.,
Rodriguez-Cintron W.,
Rodriguez-Santana J. R.,
Borrell L. N.,
LemanskeR. F.Jr., Bleecker E. R.,
Meyers D. A.,
London S. J.,
Barnes K. C.,
Raby B. A.,
Martinez F. D.,
Gilliland F. D.,
Williams L. K.,
Burchard E. G.,
Weiss S. T.,
Nicolae D. L., and
Ober C., Further replication studies of the EVE Consortium meta-analysis identifies 2 asthma risk loci in European Americans, The Journal of Allergy and Clinical Immunology. (2012) 130, no. 6, 1294–1301, https://doi.org/10.1016/j.jaci.2012.07.054, 2-s2.0-84874444273.
49Xie H.,
Li L.,
Xiong L.-D.,
Liao F., and
Zhang G.-R., The changes of skin barrier of patients with different facial dermatitis and the comparison of CE and KLK5, Journal of Sichuan University. (2013) 44, no. 6, 940–998, 2-s2.0-84893375713.
50Fortugno P.,
Furio L.,
Teson M.,
Berretti M.,
El Hachem M.,
Zambruno G.,
Hovnanian A., and
D′Alessio M., The 420k LEKTI variant alters LEKTI proteolytic activation and results in protease deregulation: implications for atopic dermatitis, Human Molecular Genetics. (2012) 21, no. 19, 4187–4200, https://doi.org/10.1093/hmg/dds243, 2-s2.0-84866353711.
51Izumi A.,
Iijima Y.,
Noguchi H.,
Numakawa T.,
Okada T.,
Hori H.,
Kato T.,
Tatsumi M.,
Kosuga A.,
Kamijima K.,
Asada T.,
Arima K.,
Saitoh O.,
Shiosaka S., and
Kunugi H., Genetic variations of human neuropsin gene and psychiatric disorders: polymorphism screening and possible association with bipolar disorder and cognitive functions, Neuropsychopharmacology. (2008) 33, no. 13, 3237–3245, https://doi.org/10.1038/npp.2008.29, 2-s2.0-38549112201.
52Li Q. D.,
Li F. J.,
Liu X. C., and
Jiang H., Klk1 A1789g gene polymorphism and the risk of coronary artery stenosis in the Chinese population, Genetics and Molecular Research. (2013) 12, no. 2, 1636–1645, https://doi.org/10.4238/2013.may.14.4, 2-s2.0-84878087175.
53Yao Y.-Y.,
Fu C.,
Ma G.-S.,
Feng Y.,
Shen C.-X.,
Wu G.-Q.,
Zhang X.-G.,
Ding J.-D.,
Tang C.-C.,
Chen Z.,
Dai Q.-M.,
Tong J.-Y.,
Luo D.,
Zhu J.,
Zhi H.,
Li Y.-J.,
Ju C.-W.,
Lu J.,
Chao J., and
Chao L., Tissue kallikrein is related to the severity of coronary artery disease, Clinica Chimica Acta. (2013) 423, 90–98, https://doi.org/10.1016/j.cca.2013.04.017, 2-s2.0-84878153729.
54Susantitaphong P.,
Perianayagam M. C.,
Kang S. W.,
Zhang W.,
Rao F.,
O′Connor D. T., and
Jaber B. L., Association of functional kallikrein-1 promoter polymorphisms and acute kidney injury: a case-control and longitudinal cohort study, Nephron. Clinical practice. (2012) 122, no. 3-4, 107–113, https://doi.org/10.1159/000350733, 2-s2.0-84893418154.
55Liu K.,
Li Q.-Z.,
Delgado-Vega A. M.,
Abelson A.-K.,
Sánchez E.,
Kelly J. A.,
Li L.,
Liu Y.,
Zhou J.,
Yan M.,
Ye Q.,
Liu S.,
Xie C.,
Zhou X. J.,
Chung S. A.,
Pons-Estel B.,
Witte T.,
De Ramón E.,
Bae S.-C.,
Barizzone N.,
Sebastiani G. D.,
Merrill J. T.,
Gregersen P. K.,
Gilkeson G. G.,
Kimberly R. P.,
Vyse T. J.,
Kim I.,
D′Alfonso S.,
Martin J.,
Harley J. B.,
Criswell L. A.,
Wakeland E. K.,
Alarcón-Riquelme M. E.,
Mohan C.,
Danieli M. G.,
Galeazzi M.,
Querini P. R.,
Migliaresi S.,
Scherbarth H. R.,
Lopez J. A.,
Motta E. L.,
Gamron S.,
Drenkard C.,
Menso E.,
Allievi A.,
Tate G. A.,
Presas J. L.,
Palatnik S. A.,
Abdala M.,
Bearzotti M.,
Alvarellos A.,
Caeiro F.,
Bertoli A.,
Paira S.,
Roverano S.,
Graf C. E.,
Bertero E.,
Caprarulo C.,
Buchanan G.,
Guillerón C.,
Grimaudo S.,
Manni J.,
Catoggio L. J.,
Soriano E. R.,
Santos C. D.,
Prigione C.,
Ramos F. A.,
Navarro S. M.,
Berbotto G. A.,
Jorfen M.,
Romero E. J.,
Garcia M. A.,
Marcos J. C.,
Marcos A. I.,
Perandones C. E.,
Eimon A.,
Battagliotti C. G.,
Armadi-Simab K.,
Gross W. L.,
Gromnica-Ihle E.,
Peter H.-H.,
Manger K.,
Schnarr S.,
Zeidler H.,
Schmidt R. E.,
Ortego N.,
Callejas J. L.,
Jiménez-Alonso J.,
Sabio M.,
Sánchez-Román J.,
Garcia-Hernandez F. J.,
Camps M.,
López-Nevot M. A.,
González-Escribano M. F.,
Langefeld C.,
Tsao B., and
Jacob C., Kallikrein genes are associated with lupus and glomerular basement membrane-specific antibody-induced nephritis in mice and humans, The Journal of Clinical Investigation. (2009) 119, no. 4, 911–923, https://doi.org/10.1172/jci36728, 2-s2.0-65249086734.
56Yiu W. H.,
Wong D. W. L.,
Chan L. Y. Y.,
Leung J. C. K.,
Chan K. W.,
Lan H. Y.,
Lai K. N., and
Tang S. C. W., Tissue kallikrein mediates pro-inflammatory pathways and activation of protease-activated receptor-4 in proximal tubular epithelial cells, PLoS ONE. (2014) 9, no. 2, e88894, https://doi.org/10.1371/journal.pone.0088894, 2-s2.0-84896112806.
57Burda J. E.,
Radulovic M.,
Yoon H., and
Scarisbrick I. A., Critical role for PAR1 in kallikrein 6-mediated oligodendrogliopathy, Glia. (2013) 61, no. 9, 1456–1470, https://doi.org/10.1002/glia.22534, 2-s2.0-84880818884.
58Scarisbrick I. A.,
Yoon H.,
Panos M.,
Larson N.,
Blaber S. I.,
Blaber M., and
Rodriguez M., Kallikrein 6 regulates early CNS demyelination in a viral model of multiple sclerosis, Brain Pathology. (2012) 22, no. 5, 709–722, https://doi.org/10.1111/j.1750-3639.2012.00577.x, 2-s2.0-84865482393.
59Hebb A. L. O.,
Bhan V.,
Wishart A. D.,
Moore C. S., and
Robertson G. S., Human kallikrein 6 cerebrospinal levels are elevated in multiple sclerosis, Current Drug Discovery Technologies. (2010) 7, no. 2, 137–140, https://doi.org/10.2174/157016310793180611, 2-s2.0-78649575448.
60Wennström M.,
Surova Y.,
Hall S.,
Nilsson C.,
Minthon L.,
Boström F.,
Hansson O., and
Nielsen H. M., Low CSF levels of both α-synuclein and the α-synuclein cleaving enzyme neurosin in patients with synucleinopathy, PLoS ONE. (2013) 8, no. 1, e53250, https://doi.org/10.1371/journal.pone.0053250, 2-s2.0-84872189834.
61Yoon H.,
Radulovic M.,
Wu J.,
Blaber S. I.,
Blaber M.,
Fehlings M. G., and
Scarisbrick I. A., Kallikrein 6 signals through PAR1 and PAR2 to promote neuron injury and exacerbate glutamate neurotoxicity, Journal of Neurochemistry. (2013) 127, no. 2, 283–298, https://doi.org/10.1111/jnc.12293, 2-s2.0-84885422046.
62Radulovic M.,
Yoon H.,
Larson N.,
Wu J.,
Linbo R.,
Burda J. E.,
Diamandis E. P.,
Blaber S. I.,
Blaber M.,
Fehlings M. G., and
Scarisbrick I. A., Kallikrein cascades in traumatic spinal cord injury: in vitro evidence for roles in axonopathy and neuron degeneration, Journal of Neuropathology and Experimental Neurology. (2013) 72, no. 11, 1072–1089, https://doi.org/10.1097/nen.0000000000000007, 2-s2.0-84891653211.
63Scarisbrick I. A.,
Radulovic M.,
Burda J. E.,
Larson N.,
Blaber S. I.,
Giannini C.,
Blaber M., and
Vandell A. G., Kallikrein 6 is a novel molecular trigger of reactive astrogliosis, Biological Chemistry. (2012) 393, no. 5, 355–367, https://doi.org/10.1515/hsz-2011-0241, 2-s2.0-84859857995.
64Rawlings A. V. and
Voegeli R., Stratum corneum proteases and dry skin conditions, Cell and Tissue Research. (2013) 351, no. 2, 217–235, https://doi.org/10.1007/s00441-012-1501-x, 2-s2.0-84891402880.
65Briot A.,
Lacroix M.,
Robin A.,
Steinhoff M.,
Deraison C., and
Hovnanian A., Par2 inactivation inhibits early production of TSLP, but not cutaneous inflammation, in netherton syndrome adult mouse model, The Journal of Investigative Dermatology. (2010) 130, no. 12, 2736–2742, https://doi.org/10.1038/jid.2010.233, 2-s2.0-78149464736.
66de Koning H. D.,
van den Bogaard E. H.,
Bergboer J. G. M.,
Kamsteeg M.,
van Vlijmen-Willems I. M. J. J.,
Hitomi K.,
Henry J.,
Simon M.,
Takashita N.,
Ishida-Yamamoto A.,
Schalkwijk J., and
Zeeuwen P. L. J. M., Expression profile of cornified envelope structural proteins and keratinocyte differentiation-regulating proteins during skin barrier repair, British Journal of Dermatology. (2012) 166, no. 6, 1245–1254, https://doi.org/10.1111/j.1365-2133.2012.10885.x, 2-s2.0-84861594725.
67Kanitakis J.,
Lora V.,
Chouvet B.,
Zambruno G.,
Haftek M., and
Faure M., Circumscribed palmo-plantar hypokeratosis: a disease of desquamation? Immunohistological study of five cases and literature review, Journal of the European Academy of Dermatology and Venereology. (2011) 25, no. 3, 296–301, https://doi.org/10.1111/j.1468-3083.2010.03784.x, 2-s2.0-79751470408.
68Shingaki K.,
Matsuzaki S.,
Taniguchi M.,
Kubo T.,
Fujiwara T.,
Kanazawa S.,
Yamamoto A.,
Tamura H.,
Maeda T.,
Ooi K.,
Matsumoto K.,
Shiosaka S., and
Tohyama M., Molecular mechanism of kallikrein-related peptidase 8/neuropsin-induced hyperkeratosis in inflamed skin, British Journal of Dermatology. (2010) 163, no. 3, 466–475, https://doi.org/10.1111/j.1365-2133.2010.09864.x, 2-s2.0-77955861149.
69Shingaki K.,
Taniguchi M.,
Kanazawa S.,
Matsuzaki S.,
Maeda T.,
Miyata S.,
Kubo T.,
Torii K.,
Shiosaka S., and
Tohyama M., NGF-p75 and neuropsin/KLK8 pathways stimulate each other to cause hyperkeratosis and acanthosis in inflamed skin, Journal of Dermatological Science. (2012) 67, no. 1, 71–73, https://doi.org/10.1016/j.jdermsci.2012.03.008, 2-s2.0-84862192539.
70Hovnanian A., Netherton syndrome: skin inflammation and allergy by loss of protease inhibition, Cell and Tissue Research. (2013) 351, no. 2, 289–300, https://doi.org/10.1007/s00441-013-1558-1, 2-s2.0-84891415326.
71Furio L.,
de Veer S.,
Jaillet M.,
Briot A.,
Robin A.,
Deraison C., and
Hovnanian A., Transgenic kallikrein 5 mice reproduce major cutaneous and systemic hallmarks of Netherton syndrome, Journal of Experimental Medicine. (2014) 211, no. 3, 499–513, https://doi.org/10.1084/jem.20131797, 2-s2.0-84896899981.
72Briot A.,
Deraison C.,
Lacroix M.,
Bonnart C.,
Robin A.,
Besson C.,
Dubus P., and
Hovnanian A., Kallikrein 5 induces atopic dermatitis-like lesions through PAR2-mediated thymic stromal lymphopoietin expression in Netherton syndrome, Journal of Experimental Medicine. (2009) 206, no. 5, 1135–1147, https://doi.org/10.1084/jem.20082242, 2-s2.0-66049125101.
73Yamasaki K. and
Gallo R. L., Rosacea as a disease of cathelicidins and skin innate immunity, Journal of Investigative Dermatology Symposium Proceedings. (2011) 15, no. 1, 12–15, https://doi.org/10.1038/jidsymp.2011.4, 2-s2.0-80955136390.
74Eissa A.,
Cretu D.,
Soosaipillai A.,
Thavaneswaran A.,
Pellett F.,
Diamandis A.,
Cevikbas F.,
Steinhoff M.,
Diamandis E. P.,
Gladman D., and
Chandran V., Serum kallikrein-8 correlates with skin activity, but not psoriatic arthritis, in patients with psoriatic disease, Clinical Chemistry and Laboratory Medicine. (2013) 51, no. 2, 317–325, https://doi.org/10.1515/cclm-2012-0251, 2-s2.0-84875436532.
75Komatsu N.,
Saijoh K.,
Kuk C.,
Shirasaki F.,
Takehara K., and
Diamandis E. P., Aberrant human tissue kallikrein levels in the stratum corneum and serum of patients with psoriasis: dependence on phenotype, severity and therapy, British Journal of Dermatology. (2007) 156, no. 5, 875–883, https://doi.org/10.1111/j.1365-2133.2006.07743.x, 2-s2.0-34247350818.
76Ximerakis M.,
Pampalakis G.,
Roumeliotis T. I.,
Sykioti V.-S.,
Garbis S. D.,
Stefanis L.,
Sotiropoulou G., and
Vekrellis K., Resistance of naturally secreted α-synuclein to proteolysis, The FASEB Journal. (2014) 28, no. 7, 3146–3158, https://doi.org/10.1096/fj.13-245852, 2-s2.0-84903714886.
77El Annan J.,
Jiang G.,
Wang D.,
Zhou J.,
Foulks G. N., and
Shao H., Elevated immunoglobulin to tissue KLK11 in patients with Sjögren syndrome, Cornea. (2013) 32, no. 5, e90–e93, https://doi.org/10.1097/ico.0b013e31826a1e2e, 2-s2.0-84876285654.
78Avgeris M.,
Mavridis K., and
Scorilas A., Kallikrein-related peptidases in prostate, breast, and ovarian cancers: from pathobiology to clinical relevance, Biological Chemistry. (2012) 393, no. 5, 301–317, https://doi.org/10.1515/hsz-2011-0260, 2-s2.0-84859810060.
79Kioulafa M.,
Kaklamanis L.,
Stathopoulos E.,
Mavroudis D.,
Georgoulias V., and
Lianidou E. S., Kallikrein 10 (KLK10) methylation as a novel prognostic biomarker in early breast cancer, Annals of Oncology. (2009) 20, no. 6, 1020–1025, https://doi.org/10.1093/annonc/mdn733, 2-s2.0-66149102406.
80Pampalakis G.,
Prosnikli E.,
Agalioti T.,
Vlahou A.,
Zoumpourlis V., and
Sotiropoulou G., A tumor-protective role for Human kallikrein-related peptidase 6 in breast cancer mediated by inhibition of epithelial-to-mesenchymal transition, Cancer Research. (2009) 69, no. 9, 3779–3787, https://doi.org/10.1158/0008-5472.CAN-08-1976, 2-s2.0-65949090534.
81Talieri M.,
Devetzi M.,
Scorilas A.,
Pappa E.,
Tsapralis N.,
Missitzis I., and
Ardavanis A., Human kallikrein-related peptidase 12 (KLK12) splice variants expression in breast cancer and their clinical impact, Tumour Biology. (2012) 33, no. 4, 1075–1084, https://doi.org/10.1007/s13277-012-0347-x, 2-s2.0-84867606758.
82Michaelidou K.,
Ardavanis A., and
Scorilas A., Clinical relevance of the deregulated kallikrein-related peptidase 8 mRNA expression in breast cancer: a novel independent indicator of disease-free survival, Breast Cancer Research and Treatment. (2015) 152, no. 2, 323–336, https://doi.org/10.1007/s10549-015-3470-8.
83Talieri M.,
Devetzi M.,
Scorilas A.,
Prezas P.,
Ardavanis A.,
Apostolaki A., and
Karameris A., Evaluation of kallikrein-related peptidase 5 expression and its significance for breast cancer patients: association with kallikrein-related peptidase 7 expression, Anticancer Research. (2011) 31, no. 9, 3093–3100, 2-s2.0-80051637527.
84Santin A. D.,
Cane′ S.,
Bellone S.,
Bignotti E.,
Palmieri M.,
De Las Casas L. E.,
Roman J. J.,
Anfossi S.,
O′Brien T., and
Pecorelli S., The serine protease stratum corneum chymotryptic enzyme (kallikrein 7) is highly overexpressed in squamous cervical cancer cells, Gynecologic Oncology. (2004) 94, no. 2, 283–288, https://doi.org/10.1016/j.ygyno.2004.05.023, 2-s2.0-3543124742.
85Termini L.,
Maciag P. C.,
Soares F. A.,
Nonogaki S.,
Pereira S. M. M.,
Alves V. A. F.,
Longatto-Filho A., and
Villa L. L., Analysis of human kallikrein 7 expression as a potential biomarker in cervical neoplasia, International Journal of Cancer. (2010) 127, no. 2, 485–490, https://doi.org/10.1002/ijc.25046, 2-s2.0-77953716477.
86Kontos C. K.,
Chantzis D.,
Papadopoulos I. N., and
Scorilas A., Kallikrein-related peptidase 4 (KLK4) mRNA predicts short-term relapse in colorectal adenocarcinoma patients, Cancer Letters. (2013) 330, no. 1, 106–112, https://doi.org/10.1016/j.canlet.2012.11.036, 2-s2.0-84872877336.
87Christodoulou S.,
Alexopoulou D. K.,
Kontos C. K.,
Scorilas A., and
Papadopoulos I. N., Kallikrein-related peptidase-6 (KLK6) mRNA expression is an independent prognostic tissue biomarker of poor disease-free and overall survival in colorectal adenocarcinoma, Tumor Biology. (2014) 35, no. 5, 4673–4685, https://doi.org/10.1007/s13277-014-1612-y, 2-s2.0-84902963881.
88Talieri M.,
Mathioudaki K.,
Prezas P.,
Alexopoulou D. K.,
Diamandis E. P.,
Xynopoulos D.,
Ardavanis A.,
Arnogiannaki N., and
Scorilas A., Clinical significance of kallikrein-related peptidase 7 (KLK7) in colorectal cancer, Thrombosis and Haemostasis. (2009) 101, no. 4, 741–747, https://doi.org/10.1160/TH08-07-0471, 2-s2.0-62549113142.
89Alexopoulou D. K.,
Papadopoulos I. N., and
Scorilas A., Clinical significance of kallikrein-related peptidase (KLK10) mRNA expression in colorectal cancer, Clinical Biochemistry. (2013) 46, no. 15, 1453–1461, https://doi.org/10.1016/j.clinbiochem.2013.03.002, 2-s2.0-84884418596.
90Feng B.,
Xu W.-B.,
Zheng M.-H.,
Ma J.-J.,
Cai Q.,
Zhang Y.,
Ji J.,
Lu A.-G.,
Qu Y.,
Li J.-W.,
Wang M.-L.,
Hu W.-G.,
Liu B.-Y., and
Zhu Z.-G., Clinical significance of human kallikrein 10 gene expression in colorectal cancer and gastric cancer, Journal of Gastroenterology and Hepatology. (2006) 21, no. 10, 1596–1603, https://doi.org/10.1111/j.1440-1746.2006.04228.x, 2-s2.0-33747607473.
91Huang W.,
Zhong J.,
Wu L.,
Yu L.,
Tian X.,
Zhang Y., and
Li B., Downregulation and CpG island hypermethylation of NES1/hK10 gene in the pathogenesis of human gastric cancer, Cancer Letters. (2007) 251, no. 1, 78–85, https://doi.org/10.1016/j.canlet.2006.11.006.
92Jiao X.,
Lu H.-J.,
Zhai M.-M.,
Tan Z.-J.,
Zhi H.-N.,
Liu X.-M.,
Liu C.-H., and
Zhang D.-P., Overexpression of kallikrein gene 10 is a biomarker for predicting poor prognosis in gastric cancer, World Journal of Gastroenterology. (2013) 19, no. 48, 9425–9431, https://doi.org/10.3748/wjg.v19.i48.9425, 2-s2.0-84891409287.
93Devetzi M.,
Trangas T.,
Scorilas A.,
Xynopoulos D., and
Talieri M., Parallel overexpression and clinical significance of kallikrein-related peptidases 7 and 14 (KLK7KLK14) in colon cancer, Thrombosis and Haemostasis. (2013) 109, no. 4, 716–725, https://doi.org/10.1160/TH12-07-0518.
94Wen Y.-G.,
Wang Q.,
Zhou C.-Z.,
Yan D.-W.,
Qiu G.-Q.,
Yang C.,
Tang H.-M., and
Peng Z.-H., Identification and validation of Kallikrein-ralated peptidase 11 as a novel prognostic marker of gastric cancer based on immunohistochemistry, Journal of Surgical Oncology. (2011) 104, no. 5, 516–524, https://doi.org/10.1002/jso.21981, 2-s2.0-80052474106.
95Konstantoudakis G.,
Florou D.,
Mavridis K.,
Papadopoulos I. N., and
Scorilas A., Kallikrein-related peptidase 13 (KLK13) gene expressional status contributes significantly in the prognosis of primary gastric carcinomas, Clinical Biochemistry. (2010) 43, no. 15, 1205–1211, https://doi.org/10.1016/j.clinbiochem.2010.07.016, 2-s2.0-77956714174.
96Worsham M. J.,
Chen K. M.,
Meduri V.,
Nygren A. O. H.,
Errami A.,
Schouten J. P., and
Benninger M. S., Epigenetic events of disease progression in head and neck squamous cell carcinoma, Archives of Otolaryngology—Head and Neck Surgery. (2006) 132, no. 6, 668–677, https://doi.org/10.1001/archotol.132.6.668, 2-s2.0-33745227364.
97Pettus J. R.,
Johnson J. J.,
Shi Z.,
Davis J. W.,
Koblinski J.,
Ghosh S.,
Liu Y.,
Ravosa M. J.,
Frazier S., and
Stack M. S., Multiple kallikrein (KLK 5, 7, 8, and 10) expression in squamous cell carcinoma of the oral cavity, Histology and Histopathology. (2009) 24, no. 2, 197–207, 2-s2.0-59149105913.
98Liu C.-J.,
Liu T.-Y.,
Kuo L.-T.,
Cheng H.-W.,
Chu T.-H.,
Chang K.-W., and
Lin S.-C., Differential gene expression signature between primary and metastatic head and neck squamous cell carcinoma, The Journal of Pathology. (2008) 214, no. 4, 489–497, https://doi.org/10.1002/path.2306, 2-s2.0-40349103687.
99Zhao H.,
Dong Y.,
Quan J.,
Smith R.,
Lam A.,
Weinstein S.,
Clements J.,
Johnson N. W., and
Gao J., Correlation of the expression of human kallikrein-related peptidases 4 and 7 with the prognosis in oral squamous cell carcinoma, Head and Neck. (2011) 33, no. 4, 566–572, https://doi.org/10.1002/hed.21496, 2-s2.0-79952923204.
100Talieri M.,
Zoma M.,
Devetzi M.,
Scorilas A., and
Ardavanis A., Kallikrein-related peptidase 6 (KLK6)gene expression in intracranial tumors, Tumor Biology. (2012) 33, no. 5, 1375–1383, https://doi.org/10.1007/s13277-012-0385-4, 2-s2.0-84868147958.
101Prezas P.,
Scorilas A.,
Yfanti C.,
Viktorov P.,
Agnanti N.,
Diamandis E., and
Talieri M., The role of human tissue kallikreins 7 and 8 in intracranial malignancies, Biological Chemistry. (2006) 387, no. 12, 1607–1612, https://doi.org/10.1515/BC.2006.200, 2-s2.0-33751549068.
102Planque C.,
Li L.,
Zheng Y.,
Soosaipillai A.,
Reckamp K.,
Chia D.,
Diamandis E. P., and
Goodglick L., A multiparametric serum kallikrein panel for diagnosis of non-small cell lung carcinoma, Clinical Cancer Research. (2008) 14, no. 5, 1355–1362, https://doi.org/10.1158/1078-0432.ccr-07-4117, 2-s2.0-40949162278.
104Singh J.,
Naran A.,
Misso N. L.,
Rigby P. J.,
Thompson P. J., and
Bhoola K. D., Expression of kallikrein-related peptidases (KRP/hK5, 7, 6, 8) in subtypes of human lung carcinoma, International Immunopharmacology. (2008) 8, no. 2, 300–306, https://doi.org/10.1016/j.intimp.2007.08.015, 2-s2.0-37549037823.
105Sher Y.-P.,
Chou C.-C.,
Chou R.-H.,
Wu H.-M.,
Chang W.-S. W.,
Chen C.-H.,
Yang P.-C.,
Wu C.-W.,
Yu C.-L., and
Peck K., Human kallikrein 8 protease confers a favorable clinical outcome in non-small cell lung cancer by suppressing tumor cell invasiveness, Cancer Research. (2006) 66, no. 24, 11763–11770, https://doi.org/10.1158/0008-5472.CAN-06-3165, 2-s2.0-33846241278.
106Zhang Y.,
Wang R.,
Song H.,
Huang G.,
Yi J.,
Zheng Y.,
Wang J., and
Chen L., Methylation of multiple genes as a candidate biomarker in non-small cell lung cancer, Cancer Letters. (2011) 303, no. 1, 21–28, https://doi.org/10.1016/j.canlet.2010.12.011, 2-s2.0-79951850937.
107Krenzer S.,
Peterziel H.,
Mauch C.,
Blaber S. I.,
Blaber M.,
Angel P., and
Hess J., Expression and function of the kallikrein-related peptidase 6 in the human melanoma microenvironment, Journal of Investigative Dermatology. (2011) 131, no. 11, 2281–2288, https://doi.org/10.1038/jid.2011.190, 2-s2.0-80054757887.
108Martins W. K.,
Esteves G. H.,
Almeida O. M.,
Rezze G. G.,
Landman G.,
Marques S. M.,
Carvalho A. F.,
L Reis L. F.,
Duprat J. P., and
Stolf B. S., Gene network analyses point to the importance of human tissue kallikreins in melanoma progression, BMC Medical Genomics. (2011) 4, article 76, https://doi.org/10.1186/1755-8794-4-76, 2-s2.0-80054931151.
109Oikonomopoulou K.,
Scorilas A.,
Michael I. P.,
Grass L.,
Soosaipillai A.,
Rosen B.,
Murphy J., and
Diamandis E. P., Kallikreins as markers of disseminated tumour cells in ovarian cancer—a pilot study, Tumor Biology. (2006) 27, no. 2, 104–114, https://doi.org/10.1159/000092325, 2-s2.0-33646118004.
110Loessner D.,
Quent V. M. C.,
Kraemer J.,
Weber E. C.,
Hutmacher D. W.,
Magdolen V., and
Clements J. A., Combined expression of KLK4, KLK5, KLK6, and KLK7 by ovarian cancer cells leads to decreased adhesion and paclitaxel-induced chemoresistance, Gynecologic Oncology. (2012) 127, no. 3, 569–578, https://doi.org/10.1016/j.ygyno.2012.09.001, 2-s2.0-84868565571.
111Dorn J.,
Harbeck N.,
Kates R.,
Gkazepis A.,
Scorilas A.,
Soosaipillai A.,
Diamandis E.,
Kiechle M.,
Schmalfeldt B., and
Schmitt M., Impact of expression differences of kallikrein-related peptidases and of uPA and PAI-1 between primary tumor and omentum metastasis in advanced ovarian cancer, Annals of Oncology. (2011) 22, no. 4, 877–883, https://doi.org/10.1093/annonc/mdq462, 2-s2.0-79953309892.
113Bandiera E.,
Zanotti L.,
Fabricio A. S. C.,
Bucca E.,
Squarcina E.,
Romani C.,
Tassi R.,
Bignotti E.,
Todeschini P.,
Tognon G.,
Romagnolo C.,
Gion M.,
Sartori E.,
Maggino T.,
Pecorelli S., and
Ravaggi A., Cancer antigen 125, human epididymis 4, kallikrein 6, osteopontin and soluble mesothelin-related peptide immunocomplexed with immunoglobulin M in epithelial ovarian cancer diagnosis, Clinical Chemistry and Laboratory Medicine. (2013) 51, no. 9, 1815–1824, https://doi.org/10.1515/cclm-2013-0151, 2-s2.0-84888097593.
114Sarojini S.,
Tamir A.,
Lim H.,
Li S.,
Zhang S.,
Goy A.,
Pecora A., and
Suh K. S., Early detection biomarkers for ovarian cancer, Journal of Oncology. (2012) 2012, 15, 709049, https://doi.org/10.1155/2012/709049, 2-s2.0-84872790574.
115Koh S. C. L.,
Razvi K.,
Chan Y. H.,
Narasimhan K.,
Ilancheran A.,
Low J. J.,
Choolani M.,
Sahil F.,
Lutan D.,
Siddik D.,
Suwiyoga K.,
Budiana G.,
Pradjitmo H.,
Saleh A. Z.,
Aziz M. F.,
Winarto H.,
Thanh P. V.,
Hoan N. K. H.,
Ravichandran I., and
Dominic N., The association with age, human tissue kallikreins 6 and 10 and hemostatic markers for survival outcome from epithelial ovarian cancer, Archives of Gynecology and Obstetrics. (2011) 284, no. 1, 183–190, https://doi.org/10.1007/s00404-010-1605-z, 2-s2.0-79958793918.
116Gunawardana C. G.,
Kuk C.,
Smith C. R.,
Batruch I.,
Soosaipillai A., and
Diamandis E. P., Comprehensive analysis of conditioned media from ovarian cancer cell lines identifies novel candidate markers of epithelial ovarian cancer, Journal of Proteome Research. (2009) 8, no. 10, 4705–4713, https://doi.org/10.1021/pr900411g, 2-s2.0-70349974679.
117Psyrri A.,
Kountourakis P.,
Scorilas A.,
Markakis S.,
Camp R.,
Kowalski D.,
Diamandis E. P., and
Dimopoulos M. A., Human tissue kallikrein 7, a novel biomarker for advanced ovarian carcinoma using a novel in situ quantitative method of protein expression, Annals of Oncology. (2008) 19, no. 7, 1271–1277, https://doi.org/10.1093/annonc/mdn035, 2-s2.0-46249126804.
118Kountourakis P.,
Psyrri A.,
Scorilas A.,
Markakis S.,
Kowaiski D.,
Camp R. L.,
Diamandis E. P., and
Dimopoulos M. A., Expression and prognostic significance of kallikrein-related peptidase 8 protein levels in advanced ovarian cancer by using automated quantitative analysis, Thrombosis and Haemostasis. (2009) 101, no. 3, 541–546, https://doi.org/10.1160/TH08-01-0052, 2-s2.0-62549103358.
119Borgoño C. A.,
Kishi T.,
Scorilas A.,
Harbeck N.,
Dorn J.,
Schmalfeldt B.,
Schmitt M., and
Diamandis E. P., Human kallikrein 8 protein is a favorable prognostic marker in ovarian cancer, Clinical Cancer Research. (2006) 12, no. 5, 1487–1493, https://doi.org/10.1158/1078-0432.ccr-05-2106, 2-s2.0-33645094382.
120Yousef G. M.,
Kyriakopoulou L. G.,
Scorilas A.,
Fracchioli S.,
Ghiringhello B.,
Zarghooni M.,
Chang A.,
Diamandis M.,
Giardina G.,
Hartwick W. J.,
Richiardi G.,
Massobrio M.,
Diamandis E. P., and
Katsaros D., Quantitative expression of the human kallikrein gene 9 (KLK9) in ovarian cancer: a new independent and favorable prognostic marker, Cancer Research. (2001) 61, no. 21, 7811–7818, 2-s2.0-0035502995.
121Batra J.,
Tan O. L.,
O′Mara T.,
Zammit R.,
Nagle C. M.,
Clements J. A.,
Kedda M.-A., and
Spurdle A. B., Kallikrein-related peptidase 10 (KLK10) expression and single nucleotide polymorphisms in ovarian cancer survival, International Journal of Gynecological Cancer. (2010) 20, no. 4, 529–536, https://doi.org/10.1111/igc.0b013e3181d9273e, 2-s2.0-77952246142.
122Borgoño C. A.,
Fracchiol S.,
Yousef G. M.,
Rigault de la Longrais I. A.,
Luo L.-Y.,
Soosaipillai A.,
Puopolo M.,
Grass L.,
Scorilas A.,
Diamandis E. P., and
Katsaros D., Favorable prognostic value of tissue human kallikrein 11 (hK11) in patients with ovarian carcinoma, International Journal of Cancer. (2003) 106, no. 4, 605–610, https://doi.org/10.1002/ijc.11296, 2-s2.0-0042594383.
123Scorilas A.,
Borgono C. A.,
Harbeck N.,
Dorn J.,
Schmalfeldt B.,
Schmitt M., and
Diamandis E. P., Human kallikrein 13 protein in ovarian cancer cytosols: a new favorable prognostic marker, Journal of Clinical Oncology. (2004) 22, 678–685.
124Yousef G. M.,
Fracchioli S.,
Scorilas A.,
Borgoño C. A.,
Iskander L.,
Puopolo M.,
Massobrio M.,
Diamandis E. P., and
Katsaros D., Steroid hormone regulation and prognostic value of the human kallikrein gene 14 in ovarian cancer, American Journal of Clinical Pathology. (2003) 119, no. 3, 346–355, https://doi.org/10.1309/0ua57mnayv0mce9u, 2-s2.0-0037364730.
125Batra J.,
Nagle C. M.,
O′Mara T.,
Higgins M.,
Dong Y.,
Tan O. L.,
Lose F.,
Skeie L. M.,
Srinivasan S.,
Bolton K. L.,
Song H.,
Ramus S. J.,
Gayther S. A.,
Pharoah P. D. P.,
Kedda M.-A.,
Spurdle A. B., and
Clements J. A., A Kallikrein 15 (KLK15) single nucleotide polymorphism located close to a novel exon shows evidence of association with poor ovarian cancer survival, BMC Cancer. (2011) 11, article 119, https://doi.org/10.1186/1471-2407-11-119, 2-s2.0-79953253977.
126Iakovlev V.,
Siegel E. R.,
Tsao M.-S., and
Haun R. S., Expression of kallikrein-related peptidase 7 predicts poor prognosis in patients with unresectable pancreatic ductal adenocarcinoma, Cancer Epidemiology Biomarkers and Prevention. (2012) 21, no. 7, 1135–1142, https://doi.org/10.1158/1055-9965.EPI-11-1079, 2-s2.0-84863588450.
127Lawrence M. G.,
Lai J., and
Clements J. A., Kallikreins on steroids: structure, function, and hormonal regulation of prostate-specific antigen and the extended kallikrein locus, Endocrine Reviews. (2010) 31, no. 4, 407–446, https://doi.org/10.1210/er.2009-0034, 2-s2.0-77955373045.
128Veveris-Lowe T. L.,
Kruger S. J.,
Walsh T.,
Gardiner R. A., and
Clements J. A., Seminal fluid characterization for male fertility and prostate cancer: Kallikrein-related serine proteases and whole proteome approaches, Seminars in Thrombosis and Hemostasis. (2007) 33, no. 1, 87–99, https://doi.org/10.1055/s-2006-958467, 2-s2.0-33846615962.
129O′Malley K. J.,
Eisermann K.,
Pascal L. E.,
Parwani A. V.,
Majima T.,
Graham L.,
Hrebinko K.,
Acquafondata M.,
Stewart N. A.,
Nelson J. B.,
Yoshimura N., and
Wang Z., Proteomic analysis of patient tissue reveals PSA protein in the stroma of benign prostatic hyperplasia, Prostate. (2014) 74, no. 8, 892–900, https://doi.org/10.1002/pros.22807, 2-s2.0-84899585756.
130Thorek D. L. J.,
Evans M. J.,
Carlsson S. V.,
Ulmert D., and
Lilja H., Prostate-specific kallikrein-related peptidases and their relation to prostate cancer biology and detection, Thrombosis and Haemostasis. (2013) 110, no. 3, 484–492, https://doi.org/10.1160/TH13-04-0275, 2-s2.0-84883254543.
131Borque Á.,
del Amo J.,
Esteban L. M.,
Ars E.,
Hernández C.,
Planas J.,
Arruza A.,
Llarena R.,
Palou J.,
Herranz F.,
Raventõs C. X.,
Tejedor D.,
Artieda M.,
Simon L.,
Martínez A.,
Carceller E.,
Suárez M.,
Allué M.,
Sanz G., and
Morote J., Genetic predisposition to early recurrence in clinically localized prostate cancer, BJU International. (2013) 111, no. 4, 549–558, https://doi.org/10.1111/j.1464-410x.2012.11333.x, 2-s2.0-84875850928.
132Kohli M.,
Rothberg P. G.,
Feng C.,
Messing E.,
Joseph J.,
Rao S. S.,
Hendershot A., and
Sahsrabudhe D., Exploratory study of a KLK2 polymorphism as a prognostic marker in prostate cancer, Cancer Biomarkers. (2010) 7, no. 2, 101–108, https://doi.org/10.3233/CBM-2010-0152, 2-s2.0-78650595744.
133Nam R. K.,
Zhang W. W.,
Klotz L. H.,
Trachtenberg J.,
Jewett M. A. S.,
Sweet J.,
Toi A.,
Teahan S.,
Venkateswaran V.,
Sugar L.,
Loblaw A.,
Siminovitch K., and
Narod S. A., Variants of the hK2 protein gene (KLK2) are associated with serum hK2 levels and predict the presence of prostate cancer at biopsy, Clinical Cancer Research. (2006) 12, no. 21, 6452–6458, https://doi.org/10.1158/1078-0432.CCR-06-1485, 2-s2.0-33751261645.
134Amaro A.,
Esposito A. I.,
Gallina A.,
Nees M.,
Angelini G.,
Albini A., and
Pfeffer U., Validation of proposed prostate cancer biomarkers with gene expression data: a long road to travel, Cancer and Metastasis Reviews. (2014) 33, no. 2, 657–671, https://doi.org/10.1007/s10555-013-9470-4, 2-s2.0-84905092743.
135Shui I. M.,
Lindström S.,
Kibel A. S.,
Berndt S. I.,
Campa D.,
Gerke T.,
Penney K. L.,
Albanes D.,
Berg C.,
Bueno-De-Mesquita H. B.,
Chanock S.,
Crawford E. D.,
Diver W. R.,
Gapstur S. M.,
Gaziano J. M.,
Giles G. G.,
Henderson B.,
Hoover R.,
Johansson M.,
Le Marchand L.,
Ma J.,
Navarro C.,
Overvad K.,
Schumacher F. R.,
Severi G.,
Siddiq A.,
Stampfer M.,
Stevens V. L.,
Travis R. C.,
Trichopoulos D.,
Vineis P.,
Mucci L. A.,
Yeager M.,
Giovannucci E., and
Kraft P., Prostate cancer (PCa) risk variants and risk of fatal PCa in the national cancer institute breast and prostate cancer cohort consortium, European Urology. (2014) 65, no. 6, 1069–1075, https://doi.org/10.1016/j.eururo.2013.12.058, 2-s2.0-84899529644.
136Soni A.,
Bansal A.,
Mishra A. K.,
Batra J.,
Singh L. C.,
Chakraborty A.,
Yadav D. S.,
Mohanty N. K., and
Saxena S., Association of androgen receptor, prostate-specific antigen, and CYP19 gene polymorphisms with prostate carcinoma and benign prostatic hyperplasia in a north Indian population, Genetic Testing and Molecular Biomarkers. (2012) 16, no. 8, 835–840, https://doi.org/10.1089/gtmb.2011.0322, 2-s2.0-84865675188.
137Kwon E. M.,
Holt S. K.,
Fu R.,
Kolb S.,
Williams G.,
Stanford J. L., and
Ostrander E. A., Androgen metabolism and JAK/STAT pathway genes and prostate cancer risk, Cancer Epidemiology. (2012) 36, no. 4, 347–353, https://doi.org/10.1016/j.canep.2012.04.002, 2-s2.0-84863436368.
138Nobata S.,
Hishida A.,
Naito M.,
Asai Y.,
Mori A.,
Kuwabara M.,
Katase S.,
Okada R.,
Morita E.,
Kawai S.,
Hamajima N., and
Wakai K., Association between KLK3 rs2735839 G/A polymorphism and serum psa levels in Japanese men, Urologia Internationalis. (2012) 89, no. 1, 39–44, https://doi.org/10.1159/000332197, 2-s2.0-84864608995.
139Penney K. L.,
Schumacher F. R.,
Kraft P.,
Mucci L. A.,
Sesso H. D.,
Ma J.,
Niu Y.,
Cheong J. K.,
Hunter D. J.,
Stampfer M. J., and
Hsu S. I., Association of KLK3 (PSA) genetic variants with prostate cancer risk and PSA levels, Carcinogenesis. (2011) 32, no. 6, 853–859, https://doi.org/10.1093/carcin/bgr050, 2-s2.0-79958227889.
140Mavridis K. and
Scorilas A., Prognostic value and biological role of the kallikrein-related peptidases in human malignancies, Future Oncology. (2010) 6, no. 2, 269–285, https://doi.org/10.2217/fon.09.149, 2-s2.0-77952487860.
141Mavridis K.,
Stravodimos K., and
Scorilas A., Quantified KLK15 gene expression levels discriminate prostate cancer from benign tumors and constitute a novel independent predictor of disease progression, Prostate. (2013) 73, no. 11, 1191–1201, https://doi.org/10.1002/pros.22667, 2-s2.0-84879359041.
142Lose F.,
Srinivasan S.,
O′Mara T.,
Marquart L.,
Chambers S.,
Gardiner R. A.,
Aitken J. F.,
Yeadon T.,
Saunders P.,
Eckert A.,
Clements J.,
Heathcote P.,
Wood G.,
Malone G.,
Samaratunga H.,
Collins A.,
Turner M.,
Kerr K.,
Spurdle A. B.,
Batra J., and
Clements J. A., Genetic association of the KLK4 locus with risk of prostate cancer, PLoS ONE. (2012) 7, no. 9, e44520, https://doi.org/10.1371/journal.pone.0044520, 2-s2.0-84866060232.
143Lose F.,
Batra J.,
O′Mara T.,
Fahey P.,
Marquart L.,
Eeles R. A.,
Easton D. F.,
Al Olama A. A.,
Kote-Jarai Z.,
Guy M.,
Muir K.,
Lophatananon A.,
Rahman A. A.,
Neal D. E.,
Hamdy F. C.,
Donovan J. L.,
Chambers S.,
Gardiner R. A.,
Aitken J. F.,
Yaxley J.,
Alexander K.,
Clements J. A.,
Spurdle A. B., and
Kedda M.-A., Common variation in Kallikrein genes KLK5, KLK6, KLK12, and KLK13 and risk of prostate cancer and tumor aggressiveness, Urologic Oncology: Seminars and Original Investigations. (2013) 31, no. 5, 635–643, https://doi.org/10.1016/j.urolonc.2011.05.011, 2-s2.0-84879462937.
144Mo L.,
Zhang J.,
Shi J.,
Xuan Q.,
Yang X.,
Qin M.,
Lee C.,
Klocker H.,
Li Q. Q., and
Mo Z., Human kallikrein 7 induces epithelial-mesenchymal transition-like changes in prostate carcinoma cells: a role in prostate cancer invasion and progression, Anticancer Research. (2010) 30, no. 9, 3413–3420, 2-s2.0-77958566329.
145Bi X.,
He H.,
Ye Y.,
Dai Q.,
Han Z.,
Liang Y., and
Zhong W., Association of TMPRSS2 and KLK11 gene expression levels with clinical progression of human prostate cancer, Medical Oncology. (2010) 27, no. 1, 145–151, https://doi.org/10.1007/s12032-009-9185-0, 2-s2.0-77949266077.
146Lose F.,
Lawrence M. G.,
Srinivasan S.,
O′Mara T.,
Marquart L.,
Chambers S.,
Gardiner R. A.,
Aitken J. F.,
Spurdle A. B.,
Batra J., and
Clements J. A., The kallikrein 14 gene is down-regulated by androgen receptor signalling and harbours genetic variation that is associated with prostate tumour aggressiveness, Biological Chemistry. (2012) 393, no. 5, 403–412, https://doi.org/10.1515/hsz-2011-0268, 2-s2.0-84859843770.
147Kaushal A.,
Myers S. A.,
Dong Y.,
Lai J.,
Tan O. L.,
Bui L. T.,
Hunt M. L.,
Digby M. R.,
Samaratunga H.,
Gardiner R. A.,
Clements J. A., and
Hooper J. D., A novel transcript from the KLKP1 gene is androgen regulated, down-regulated during prostate cancer progression and encodes the first non-serine protease identified from the human kallikrein gene locus, Prostate. (2008) 68, no. 4, 381–399, https://doi.org/10.1002/pros.20685, 2-s2.0-40349088828.
149Yousef G. M.,
Chang A.,
Scorilas A., and
Diamandis E. P., Genomic organization of the human kallikrein gene family on chromosome 19q13.3-q13.4, Biochemical and Biophysical Research Communications. (2000) 276, no. 1, 125–133, https://doi.org/10.1006/bbrc.2000.3448, 2-s2.0-0034675353.
150Yousef G. M. and
Diamandis E. P., An overview of the kallikrein gene families in humans and other species: emerging candidate tumour markers, Clinical Biochemistry. (2003) 36, no. 6, 443–452, https://doi.org/10.1016/s0009-9120(03)00055-9, 2-s2.0-0043264142.
151Yousef G. M.,
Borgono C. A.,
Michael I. P., and
Diamandis E. P., Cloning of a kallikrein pseudogene, Clinical Biochemistry. (2004) 37, no. 11, 961–967, https://doi.org/10.1016/j.clinbiochem.2004.07.012, 2-s2.0-6344220185.
152Hu J. C.-C.,
Zhang C.,
Sun X.,
Yang Y.,
Cao X.,
Ryu O., and
Simmer J. P., 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, no. 1, 1–8, https://doi.org/10.1016/s0378-1119(00)00203-1, 2-s2.0-0034643918.
153Pavlopoulou A.,
Pampalakis G.,
Michalopoulos I., and
Sotiropoulou G., Evolutionary history of tissue kallikreins, PLoS ONE. (2010) 5, no. 11, e13781, https://doi.org/10.1371/journal.pone.0013781, 2-s2.0-78249243096.
154Hasegawa M.,
Kishino H., and
Yano T.-A., Dating of the human-ape splitting by a molecular clock of mitochondrial DNA, Journal of Molecular Evolution. (1985) 22, no. 2, 160–174, https://doi.org/10.1007/BF02101694, 2-s2.0-0022376704.
155Koumandou V. L. and
Scorilas A., Evolution of the plasma and tissue kallikreins, and their alternative splicing isoforms, PLoS ONE. (2013) 8, no. 7, e68074, https://doi.org/10.1371/journal.pone.0068074, 2-s2.0-84880012432.
156Elliott M. B.,
Irwin D. M., and
Diamandis E. P., In silico identification and Bayesian phylogenetic analysis of multiple new mammalian kallikrein gene families, Genomics. (2006) 88, no. 5, 591–599, https://doi.org/10.1016/j.ygeno.2006.06.001, 2-s2.0-33749452415.
157Lundwall Å., Old genes and new genes: the evolution of the kallikrein locus, Thrombosis and Haemostasis. (2013) 110, no. 3, 469–475, https://doi.org/10.1160/th12-11-0851, 2-s2.0-84883229985.
159Son D. N.,
Li L.,
Katsuyama H.,
Komatsu N.,
Saito M.,
Tanii H., and
Saijoh K., Abundant expression of Kallikrein 1 gene in human keratinocytes was mediated by GATA3, Gene. (2009) 436, no. 1-2, 121–127, https://doi.org/10.1016/j.gene.2009.02.002, 2-s2.0-62549125370.
160Zeng M.,
Zhang Y.,
Bhat I.,
Wazer D. E.,
Band H., and
Band V., The human kallikrein 10 promoter contains a functional retinoid response element, Biological Chemistry. (2006) 387, no. 6, 741–747, https://doi.org/10.1515/bc.2006.093, 2-s2.0-33745660354.
161Shin T.,
Sumiyoshi H.,
Matsuo N.,
Satoh F.,
Nomura Y.,
Mimata H., and
Yoshioka H., Sp1 and Sp3 transcription factors upregulate the proximal promoter of the human prostate-specific antigen gene in prostate cancer cells, Archives of Biochemistry and Biophysics. (2005) 435, no. 2, 291–302, https://doi.org/10.1016/j.abb.2005.01.002, 2-s2.0-13544255499.
162Eisermann K.,
Tandon S.,
Bazarov A.,
Brett A.,
Fraizer G., and
Piontkivska H., Evolutionary conservation of zinc finger transcription factor binding sites in promoters of genes co-expressed with WT1 in prostate cancer, BMC Genomics. (2008) 9, article 337, https://doi.org/10.1186/1471-2164-9-337, 2-s2.0-49249123274.
163Tsui K.-H.,
Chang P.-L.,
Lin H.-T., and
Juang H.-H., Down-regulation of the prostate specific antigen promoter by p53 in human prostate cancer cells, Journal of Urology. (2004) 172, no. 5 I, 2035–2039, https://doi.org/10.1097/01.ju.0000138053.78518.b2, 2-s2.0-5444227752.
164Cinar B.,
Yeung F.,
Konaka H.,
Mayo M. W.,
Freeman M. R.,
Zhau H. E., and
Chung L. W. K., Identification of a negative regulatory cis-element in the enhancer core region of the prostate-specific antigen promoter: implications for intersection of androgen receptor and nuclear factor-κB signalling in prostate cancer cells, Biochemical Journal. (2004) 379, no. 2, 421–431, https://doi.org/10.1042/bj20031661, 2-s2.0-2342645462.
165Borgoño C. A.,
Michael I. P., and
Diamandis E. P., Human tissue kallikreins: physiologic roles and applications in cancer, Molecular Cancer Research. (2004) 2, no. 5, 257–280, 2-s2.0-3543150203.
166Pasic M. D.,
Olkhov E.,
Bapat B., and
Yousef G. M., Epigenetic regulation of kallikrein-related peptidases: there is a whole new world out there, Biological Chemistry. (2012) 393, no. 5, 319–330, https://doi.org/10.1515/hsz-2011-0273, 2-s2.0-84859849914.
167Roman-Gomez J.,
Jimenez-Velasco A.,
Agirre X.,
Castillejo J. A.,
Barrios M.,
Andreu E. J.,
Prosper F.,
Heiniger A., and
Torres A., 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, no. 2, 362–365, https://doi.org/10.1038/sj.leu.2403223, 2-s2.0-1242329859.
168Li B.,
Goyal J.,
Dhar S.,
Dimri G.,
Evron E.,
Sukumar S.,
Wazer D. E., and
Band V., CpG methylation as a basis for breast tumor-specific loss of NES1/kallikrein 10 expression, Cancer Research. (2001) 61, no. 21, 8014–8021, 2-s2.0-0035503704.
169Olkhov-Mitsel E.,
Van der Kwast T.,
Kron K. J.,
Ozcelik H.,
Briollais L.,
Massey C.,
Recker F.,
Kwiatkowski M.,
Fleshner N. E.,
Diamandis E. P.,
Zlotta A. R., and
Bapat B., Quantitative DNA methylation analysis of genes coding for kallikrein-related peptidases 6 and 10 as biomarkers for prostate cancer, Epigenetics. (2012) 7, no. 9, 1037–1045, https://doi.org/10.4161/epi.21524, 2-s2.0-84866090183.
170Bharaj B.,
Scorilas A.,
Diamandis E. P.,
Giai M.,
Levesque M. A.,
Sutherland D. J. A., and
Hoffman B. R., Breast cancer prognostic significance of a single nucleotide polymorphism in the proximal androgen response element of the prostate specific antigen gene promoter, Breast Cancer Research and Treatment. (2000) 61, no. 2, 111–119, https://doi.org/10.1023/a:1006459613498, 2-s2.0-0033937206.
171Song Q.,
Chao J., and
Chao L., DNA polymorphisms in the 5′-flanking region of the human tissue kallikrein gene, Human Genetics. (1997) 99, no. 6, 727–734, https://doi.org/10.1007/s004390050439, 2-s2.0-0030913970.
172Kurlender L.,
Borgono C.,
Michael I. P.,
Obiezu C.,
Elliott M. B.,
Yousef G. M., and
Diamandis E. P., A survey of alternative transcripts of human tissue kallikrein genes, Biochimica et Biophysica Acta—Reviews on Cancer. (2005) 1755, no. 1, 1–14, https://doi.org/10.1016/j.bbcan.2005.02.001, 2-s2.0-19344367163.
173Matthews B. W.,
Sigler P. B.,
Henderson R., and
Blow D. M., Three-dimensional structure of tosyl-α-chymotrypsin, Nature. (1967) 214, no. 5089, 652–656, https://doi.org/10.1038/214652a0, 2-s2.0-0014201592.
174Lesk A. M. and
Fordham W. D., Conservation and variability in the structures of serine proteinases of the chymotrypsin family, Journal of Molecular Biology. (1996) 258, no. 3, 501–537, https://doi.org/10.1006/jmbi.1996.0264, 2-s2.0-0029962944.
176Ponticelli C. and
Meroni P. L., Kallikreins and lupus nephritis, The Journal of Clinical Investigation. (2009) 119, no. 4, 768–771, https://doi.org/10.1172/jci38786, 2-s2.0-65249183081.
177Diamandis E. P.,
Yousef G. M.,
Luo L.-Y.,
Magklara A., and
Obiezu C. V., The new human kallikrein gene family: implications in carcinogenesis, Trends in Endocrinology and Metabolism. (2000) 11, no. 2, 54–60, https://doi.org/10.1016/s1043-2760(99)00225-8, 2-s2.0-0034159946.
178Shaw J. L. V. and
Diamandis E. P., Distribution of 15 human kallikreins in tissues and biological fluids, Clinical Chemistry. (2007) 53, no. 8, 1423–1432, https://doi.org/10.1373/clinchem.2007.088104, 2-s2.0-34547634419.
179Komatsu N.,
Tsai B.,
Sidiropoulos M.,
Saijoh K.,
Levesque M. A.,
Takehara K., and
Diamandis E. P., Quantification of eight tissue kallikreins in the stratum corneum and sweat, The Journal of Investigative Dermatology. (2006) 126, no. 4, 925–929, 2-s2.0-33646257790.
180Laflamme B. A. and
Wolfner M. F., Identification and function of proteolysis regulators in seminal fluid, Molecular Reproduction and Development. (2013) 80, no. 2, 80–101, https://doi.org/10.1002/mrd.22130, 2-s2.0-84873748067.
181Yoon H.,
Blaber S. I.,
Li W.,
Scarisbrick I. A., and
Blaber M., Activation profiles of human kallikrein-related peptidases by matrix metalloproteinases, Biological Chemistry. (2013) 394, no. 1, 137–147, https://doi.org/10.1515/hsz-2012-0249, 2-s2.0-84876543727.
182Serrano-Pozo A.,
Frosch M. P.,
Masliah E., and
Hyman B. T., Neuropathological alterations in Alzheimer disease, Cold Spring Harbor perspectives in medicine. (2011) 1, no. 1, a006189, https://doi.org/10.1101/cshperspect.a006189, 2-s2.0-84875484373.
183Goedert M., NEURODEGENERATION. Alzheimer′s and Parkinson′s diseases: the prion concept in relation to assembled Aβ, tau, and α-synuclein, Science. (2015) 349, no. 6248, 1255555, https://doi.org/10.1126/science.1255555.
184Mitsui S.,
Okui A.,
Uemura H.,
Mizuno T.,
Yamada T.,
Yamamura Y., and
Yamaguchi N., Decreased cerebrospinal fluid levels of neurosin (KLK6), an aging-related protease, as a possible new risk factor for Alzheimer′s disease, Annals of the New York Academy of Sciences. (2002) 977, 216–223, https://doi.org/10.1111/j.1749-6632.2002.tb04818.x, 2-s2.0-18744363878.
185Zarghooni M.,
Soosaipillai A.,
Grass L.,
Scorilas A.,
Mirazimi N., and
Diamandis E. P., Decreased concentration of human kallikrein 6 in brain extracts of Alzheimer′s disease patients, Clinical Biochemistry. (2002) 35, no. 3, 225–231, https://doi.org/10.1016/S0009-9120(02)00292-8, 2-s2.0-0036017352.
186Shropshire T. D.,
Reifert J.,
Rajagopalan S.,
Baker D.,
Feinstein S. C., and
Daugherty P. S., Amyloid β peptide cleavage by kallikrein 7 attenuates fibril growth and rescues neurons from Aβ-mediated toxicity in vitro, Biological Chemistry. (2014) 395, no. 1, 109–118, https://doi.org/10.1515/hsz-2013-0230, 2-s2.0-84890378967.
187Bayani J. and
Diamandis E. P., The physiology and pathobiology of human kallikrein-related peptidase 6 (KLK6), Clinical Chemistry and Laboratory Medicine. (2012) 50, no. 2, 211–233, https://doi.org/10.1515/cclm.2011.750, 2-s2.0-84859869422.
188Emmanouilidou E.,
Melachroinou K.,
Roumeliotis T.,
Garbis S. D.,
Ntzouni M.,
Margaritis L. H.,
Stefanis L., and
Vekrellis K., Cell-produced α-synuclein is secreted in a calcium-dependent manner by exosomes and impacts neuronal survival, The Journal of Neuroscience. (2010) 30, no. 20, 6838–6851, https://doi.org/10.1523/jneurosci.5699-09.2010, 2-s2.0-77952648551.
189Kim C.,
Ho D.-H.,
Suk J.-E.,
You S.,
Michael S.,
Kang J.,
Lee S. J.,
Masliah E.,
Hwang D.,
Lee H.-J., and
Lee S.-J., Neuron-released oligomeric α-synuclein is an endogenous agonist of TLR2 for paracrine activation of microglia, Nature Communications. (2013) 4, article 1562, https://doi.org/10.1038/ncomms2534, 2-s2.0-84875898265.
191Kasai T.,
Tokuda T.,
Yamaguchi N.,
Watanabe Y.,
Kametani F.,
Nakagawa M., and
Mizuno T., Cleavage of normal and pathological forms of α-synuclein by neurosin in vitro, Neuroscience Letters. (2008) 436, no. 1, 52–56, https://doi.org/10.1016/j.neulet.2008.02.057, 2-s2.0-41749115922.
192Scarisbrick I. A.,
Linbo R.,
Vandell A. G.,
Keegan M.,
Blaber S. I.,
Blaber M.,
Sneve D.,
Lucchinetti C. F.,
Rodriguez M., and
Diamandis E. P., Kallikreins are associated with secondary progressive multiple sclerosis and promote neurodegeneration, Biological Chemistry. (2008) 389, no. 6, 739–745, https://doi.org/10.1515/BC.2008.085, 2-s2.0-47549115421.
193Naffah-Mazzacoratti Mda G.,
Gouveia T. L.,
Simoes P. S., and
Perosa S. R., What have we learned about the kallikrein-kinin and renin-angiotensin systems in neurological disorders?, World Journal of Biological Chemistry. (2014) 5, 130–140.
194Lu Z.,
Yang Q.,
Cui M.,
Liu Y.,
Wang T.,
Zhao H., and
Dong Q., Tissue kallikrein induces SH-SY5Y cell proliferation via epidermal growth factor receptor and extracellular signal-regulated kinase1/2 pathway, Biochemical and Biophysical Research Communications. (2014) 446, no. 1, 25–29, https://doi.org/10.1016/j.bbrc.2014.02.027, 2-s2.0-84898003447.
195Jin Y.,
Qu S.,
Tesikova M.,
Wang L.,
Kristian A.,
Mælandsmo G. M.,
Kong H.,
Zhang T.,
Jerońimo C.,
Teixeira M. R.,
Yuca E.,
Tekedereli I.,
Gorgulu K.,
Alpay N.,
Sood A. K.,
Lopez-Berestein G.,
Danielsen H. E.,
Ozpolat B., and
Saatcioglu F., Molecular circuit involving KLK4 integrates androgen and mTOR signaling in prostate cancer, Proceedings of the National Academy of Sciences of the United States of America. (2013) 110, no. 28, E2572–E2581, https://doi.org/10.1073/pnas.1304318110, 2-s2.0-84879893820.
196Stone O. A.,
Richer C.,
Emanueli C.,
Van Weel V.,
Quax P. H. A.,
Katare R.,
Kraenkel N.,
Campagnolo P.,
Barcelos L. S.,
Siragusa M.,
Sala-Newby G. B.,
Baldessari D.,
Mione M.,
Vincent M. P.,
Benest A. V.,
Al Haj Zen A.,
Gonzalez J.,
Bates D. O.,
Alhenc-Gelas F., and
Madeddu P., Critical role of tissue kallikrein in vessel formation and maturation: implications for therapeutic revascularization, Arteriosclerosis, Thrombosis, and Vascular Biology. (2009) 29, no. 5, 657–664, https://doi.org/10.1161/atvbaha.108.182139, 2-s2.0-65549102213.
197Dominek P.,
Campagnolo P.,
H-Zadeh M.,
Kränkel N.,
Chilosi M.,
Sharman J. A.,
Caporali A.,
Mangialardi G.,
Spinetti G.,
Emanueli C.,
Pignatelli M., and
Madeddu P., Role of human tissue kallikrein in gastrointestinal stromal tumour invasion, British Journal of Cancer. (2010) 103, no. 9, 1422–1431, https://doi.org/10.1038/sj.bjc.6605906, 2-s2.0-78049285304.
200Hirschhaeuser F.,
Sattler U. G. A., and
Mueller-Klieser W., Lactate: a metabolic key player in cancer, Cancer Research. (2011) 71, no. 22, 6921–6925, https://doi.org/10.1158/0008-5472.can-11-1457, 2-s2.0-81155126061.
201Jiang R.,
Shi Z.,
Johnson J. J.,
Liu Y., and
Stack M. S., Kallikrein-5 promotes cleavage of desmoglein-1 and loss of cell-cell cohesion in oral squamous cell carcinoma, The Journal of Biological Chemistry. (2011) 286, no. 11, 9127–9135, https://doi.org/10.1074/jbc.m110.191361, 2-s2.0-79953230837.
202Veveris-Lowe T. L.,
Lawrence M. G.,
Collard R. L.,
Bui L.,
Herington A. C.,
Nicol D. L., and
Clements J. A., Kallikrein 4 (hK4) and prostate-specific antigen (PSA) are associated with the loss of E-cadherin and an epithelial-mesenchymal transition (EMT)-like effect in prostate cancer cells, Endocrine-Related Cancer. (2005) 12, no. 3, 631–643, https://doi.org/10.1677/erc.1.00958, 2-s2.0-26944435866.
203Prassas I.,
Eissa A.,
Poda G., and
Diamandis E. P., Unleashing the therapeutic potential of human kallikrein-related serine proteases, Nature Reviews Drug Discovery. (2015) 14, no. 3, 183–202, https://doi.org/10.1038/nrd4534.
204Sotiropoulou G. and
Pampalakis G., Targeting the kallikrein-related peptidases for drug development, Trends in Pharmacological Sciences. (2012) 33, no. 12, 623–634, https://doi.org/10.1016/j.tips.2012.09.005, 2-s2.0-84869203481.
205Tamura K.,
Peterson D.,
Peterson N.,
Stecher G.,
Nei M., and
Kumar S., MEGA5: molecular evolutionary genetics analysis using maximum likelihood, evolutionary distance, and maximum parsimony methods, Molecular Biology and Evolution. (2011) 28, no. 10, 2731–2739, https://doi.org/10.1093/molbev/msr121, 2-s2.0-79957613599.
206Felsenstein J., Confidence limits on phytogenies: an approach using the bootstrap, Evolution. (1985) 39, no. 4, 783–791, https://doi.org/10.2307/2408678.
Please check your email for instructions on resetting your password.
If you do not receive an email within 10 minutes, your email address may not be registered,
and you may need to create a new Wiley Online Library account.
Request Username
Can't sign in? Forgot your username?
Enter your email address below and we will send you your username
If the address matches an existing account you will receive an email with instructions to retrieve your username
The full text of this article hosted at iucr.org is unavailable due to technical difficulties.