Open Access

Aberrant expression and high‑frequency mutations of SHARPIN in nonmelanoma skin cancer

  • Authors:
    • Yan Zheng
    • Yao Yang
    • Jiaman Wang
    • Yanhua Liang
  • View Affiliations

  • Published online on: February 12, 2019     https://doi.org/10.3892/etm.2019.7261
  • Pages: 2746-2756
  • Copyright: © Zheng et al. This is an open access article distributed under the terms of Creative Commons Attribution License.

Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

Squamous cell carcinoma (SCC) and basal cell carcinoma (BCC) have exhibited a marked increase in incidence in previous decades and are the most common malignancies in Caucasian populations. Src homology 3 and multiple ankyrin repeat domains protein‑associated RH domain‑interacting protein (SHARPIN) has been identified as a commonly overexpressed proto‑oncogene in several types of visceral cancer. However, to the best of our knowledge, the functions of SHARPIN in nonmelanoma skin cancer (NMSC) have not been described. The present study aimed to investigate the expression of SHARPIN protein and SHARPIN mutations in NMSC. A total of 85 BCC, 77 SCC and 21 keratoacanthoma (KA) formalin‑fixed paraffin‑embedded (FFPE) samples were collected. SHARPIN expression was detected using immunohistochemistry. DNA was extracted from the FFPE samples, and the sequences of SHARPIN were analyzed using polymerase chain reaction. In addition, high and moderate expression levels of SHARPIN were observed in normal skin tissues and KA samples. However, the expression of SHARPIN was absent in cancer nests and was significantly low in precancerous NMSC lesions. The total mutation frequency of SHARPIN was 21.8% in BCC and 17.0% in SCC. These data indicate that SHARPIN may serve a tumor‑suppressing role and be a promising diagnostic, prognostic and therapeutic biomarker in NMSC.

Introduction

The incidence of nonmelanoma skin cancer (NMSC), which includes squamous cell carcinoma (SCC) and basal cell carcinoma (BCC), has exhibited a marked increase in the previous decade and at present is the most common malignancy in Caucasian populations (1). NMSC is associated with a low rate of mortality but a high rate of disfigurement in cases where skin lesions are located on the head and neck. In addition, SCC occurs less frequently compared with BCC but is generally more aggressive. Sunlight (2), viral infection (3), diet (4), immunosuppression in organ transplant recipients (5) and induction of spontaneous genetic mutations (6) have been regarded as causes for NMSC. Tumors are markedly associated with chronic ultraviolet (UV) radiation exposure and occur primarily on sun-exposed areas of the body (7). Early detection and surgical removal may prevent the majority of complications. However, skin cancer has a high rate of recurrence and occasionally tumors progress to advanced stages that are difficult to treat with present therapeutic modalities; additionally, advanced-stage tumors become associated with high morbidity and decreased survival rates (8). At present, treatment options have remained limited for locally advanced or metastatic NMSC. Therefore, an in-depth understanding of the molecular basis of skin tumorigenesis is necessary in order to develop novel and specific diagnostic biomarkers and efficient therapies.

Src homology 3 and multiple ankyrin repeat domains protein (SHANK)-associated RH domain-interacting protein (SHARPIN) is a 387-amino acid protein that was originally identified as a SHANK-binding protein, which is enriched in the postsynaptic density of excitatory neurotransmitters (9). In addition, SHARPIN has been detected in cancer in the brain, spleen, lungs and other organs. Seymour et al (10) have identified SHARPIN as a gene mutated in chronic proliferative dermatitis (cpdm) mice (Sharpincpdm/cpdm) which spontaneously causes chronic inflammation, primarily in the skin, but also in other tissues including the gut, lung, liver and esophagus. SHARPIN has been previously identified as a common component of the linear ubiquitin chain assembly complex (LUBAC) which also contains E3 ubiquitin protein ligase ring finger protein 31 and RanBP-type and C3HC4-type zinc finger-containing protein 1 (HOIL-1L) (11). The C-terminal portion of SHARPIN consists of a ubiquitin-like (UBL) domain followed by an Npl4-zinc finger (NZF) domain and is important for the formation of a complex with the LUBAC component, haem-oxidized iron-regulatory protein 2 ubiquitin ligase-1 interacting protein and ubiquitin (9). LUBAC is an important component of the nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB) signaling pathway, which is a critical regulator of inflammation, immune response and lymphoid tissue development (12). NF-κB signaling is generally classified into canonical and non-canonical pathways. The canonical pathway, primarily triggered by tumor necrosis factor (TNF), lipopolysaccharides, and T and B cell receptors, occurs in the majority of cells as the principal NF-κB pathway. Upon stimulation, the downstream kinase inhibitor of κB (IκB) kinase (IKK) complex, composed of two catalytic subunits (IKKα and IKKβ) and one regulatory subunit [NF-κB essential modulator (NEMO)], is activated, allowing the phosphorylation of the IκBα inhibitory protein. A linear form of polyubiquitin chains was previously identified in the NF-κB signaling pathway following TNF stimulation (13). The generation of linear ubiquitin polymers is catalyzed by LUBAC. Previous evidence indicates that LUBAC is recruited to TNF receptor complexes upon TNF induction, and then conjugates linear ubiquitin chains to the regulatory subunit NEMO of the IKK complex (14). This activates the kinase activity of IKK and ubiquitin-dependent degradation of phosphorylated IκBα, therefore enabling the nuclear translocation of NF-κB dimers and downstream gene expression (15). SHARPIN contains a PH (pleckstrin homology) domain at the N-terminus, which serves as a dimerization domain and may serve a role in other physiological functions of SHARPIN, including its tumor-associated role and its ability to inhibit β1-integrin activation (16). Furthermore, SHARPIN has been identified as a commonly overexpressed proto-oncogene and functionally serves tumor-associated roles during cancer progression according to previous studies (1723). However, data regarding the function of SHARPIN in the pathogenesis and development of NMSC is lacking. These background data prompted the present study to investigate the expression and mutations of SHARPIN in skin tumors and identify a promising prognostic biomarker and therapeutic target for NMSC. Immunohistochemistry was utilized in the current study to assess SHARPIN expression in NMSCs and polymerase chain reaction (PCR) was used to detect mutations of SHARPIN in NMSCs. It was revealed that the expression of SHARPIN was absent in cancer nests and was significantly low in precancerous NMSC lesions. The total mutation frequency of SHARPIN was 21.8% in BCC and 17.0% in SCC.

Materials and methods

Literature retrieval

To acquire all literature regarding SHARPIN and NMSCs, PubMed (https://www.ncbi.nlm.nih.gov/pubmed) was searched using the following search string to identify relevant papers: (NMSC) OR non-melanoma skin cancer AND SHARPIN. No restrictions on publication date or language were imposed during the search strategy. No articles were identified.

Specimen selection

Anonymized control DNA samples from blood specimens of 100 normal individuals and skin tissues from 12 healthy volunteers who received cosmetic surgeries were obtained according to a protocol approved by the Southern Medical University Shenzhen Hospital Subject Review Board. All 100 normal individuals and 12 healthy volunteers did not have skin diseases. Formalin-fixed paraffin-embedded (FFPE) samples were retrieved from the Department of Dermatology of Shenzhen Hospital in Southern Medical University (Shenzhen, China). All samples from January 2012 to June 2017 were biopsied. All samples were fixed for 24 h in 10% formalin solution at room temperature. The thickness of the sections was 4 µm. A total of 85 BCC, 77 SCC and 21 keratoacanthoma (KA) FFPE samples were collected. The diagnoses of the samples were confirmed by pathologists from the Department of Dermatology of Shenzhen Hospital in Southern Medical University. Informed consent was obtained from all patients.

DNA extraction and mutation sequencing

DNA was extracted from the blood using the phenol-chloroform method (24). The FFPE genomic DNA was extracted using a QIAamp DNA FFPE Tissue kit (Qiagen GmbH, Hilden, Germany). To detect hotspot mutations, 8 exons and exon-intron adjacent sequences of the SHARPIN gene were amplified using PCR. In the DNA from the tumor samples, each amplification reaction was performed under standard conditions in a 20 µl PCR mixture containing 70–150 ng template DNA, 10 pmol primers, and 10 µl 2X Taq Master Mix (Dye Plus) (Vazyme, Piscataway, NJ, USA). The GC percentage of Exon 1 was relatively high; therefore, the 2X Taq Master Mix (Dye Plus) was replaced by 2X Phanta Max Master Mix (Vazyme) in the amplification of Exon 1. The 8 primer pairs that were used are listed in Table I. Exon 3 was amplified by PCR. The thermocycler conditions for the standard and nested PCR protocols are listed in Table II. PCR products were purified using QIAquick reagent (Qiagen GmbH) and directly sequenced based on the Big Dye Terminator sequencing chemistry (Applied Biosystems; Thermo Fisher Scientific, Inc., Waltham, MA USA) in an ABI3130 automated sequencer (Applied Biosystems; Thermo Fisher Scientific, Inc.). All mutations were confirmed through repeated bidirectional sequencing on the ABI sequencer. Gene sequences were blasted using DNASTAR Lasergene 7.1 (DNASTAR Inc., Madison, WI, USA).

Table I.

Primers used in the screening of Src homology 3 and multiple ankyrin repeat domains protein-associated RH domain-interacting protein gene mutations.

Table I.

Primers used in the screening of Src homology 3 and multiple ankyrin repeat domains protein-associated RH domain-interacting protein gene mutations.

ExonForward primer (5′-3′)Reverse primer (5′-3′)
1 CAGGTTCGCGGCCCGTGTTT AAGAGGACTGACCGCGCGCC
2 ATTTCTTTGCTCCTCGTGCG CTTCCCAGACATCCAGCAGT
3 CAGCACAGCACACCCATATC GGGACTATCTGCTATCCCCG
4 AGCAGATAGTCCCCAGTGGT GTGGGTTCAGGGATGGATGG
5 CATCAGGTGAGGCCTGGG CCGAGCTCTGAGAACACCTG
6 ATCACCTGCCCTGATGCTC GTGGAGCTCAGGACTGTGG
7 CACAGTCCTGAGCTCCACC GTTGCTTCCCTGCTCTTTCC
8 CAGGGAAGCAACAACTGTCT ATTCCTGTGGATTCTGCCCT

Table II.

PCR amplification thermocycler conditions of Src homology 3 and multiple ankyrin repeat domains protein-associated RH domain-interacting protein gene.

Table II.

PCR amplification thermocycler conditions of Src homology 3 and multiple ankyrin repeat domains protein-associated RH domain-interacting protein gene.

Touchdown PCROrdinary PCR


StepsTemperature, °CDurationStepsTemperature, °CDuration
    1945 min1945 min
    29430 sec29430 sec
    363, EXa −0.530 sec360/56/57b30 sec
    47220 sec47220 sec
    5Back to step 216 times5Back to step 235 times
    69430 sec6727 min
    75430 sec7  4Until use
    87220 sec
    9Back to step 520 times
10727 min
11  4Until use

[i] PCR, polymerase chain reaction; E, exon. E3 was amplified by Touchdown PCR. E1, E2, E4, E5, E6, E7 and E8 were amplified by ordinary PCR. aEX indicates that the annealing temperature decreased by 0.5°C per cycle. bAnnealing temperature of E1 was 60°C, annealing temperature of E2, E4, E5, and E8 was 56°C and annealing temperature of E6 and E7 was 57°C.

Immunohistochemistry

FFPE sections were deparaffinized in xylene at room temperature and rehydrated in 100, 95, 90, 80 and 70% alcohol solutions prepared with absolute ethyl alcohol and distilled water. For antigen retrieval, sections were heated in citrate buffer (pH 6.0) for 15 min at 100°C in a microwave oven and naturally cooled to room temperature. Subsequently, the samples were blocked with a mixture of methanol and 0.75% hydrogen peroxide for 20 min at room temperature. Following washing with PBS, samples were incubated with SHARPIN antibody (cat. no., sc-98127; Santa Cruz Biotechnology, Inc., Dallas, TX, USA; dilution, 1:100) at 4°C overnight. Subsequent to incubation, slides were washed three times with PBS. The slides were then processed using a 2-step Plus® Poly-horseradish peroxidase Anti-Mouse/Rabbit IgG Detection System (cat. no., PV-9000; ZSGB-BIO; OriGene Technologies, Inc., Beijing, China) and were developed with a DAB Detection kit (Enhanced Polymer; cat. no., PV-9000-D; ZSGB-BIO; OriGene Technologies, Inc.) for 3 min at room temperature. SHARPIN immunohistochemical staining was expected to be localized to the cytoplasm.

Histologic scoring and analysis

Samples were evaluated using a standard light microscopic technique (magnification, ×200) as performed by two pathologists (Shenzhen Hospital in Southern Medical University). Staining for the SHARPIN protein was evaluated in the tumors and in the normal skin tissues, which were invariably SHARPIN-positive and served as positive controls. Each tumor sample was scored by the cross-product (H score) of the percentage of tumor cell staining at each of the 3 staining intensities. Degrees of staining were divided into four levels: None, 0; weak, 1; moderate, 2; and strong, 3. For example, a particular tumor may have 30% cell staining at intensity =1 and 70% of cell staining at intensity =3, for a combined H score of 240 [(30×1) + (70×3) =240] out of a maximum of 300. This system was performed as described previously by Bollag et al (25). Concordance was observed between the scores given by the two pathologists (81% of the scores were in agreement within a 40-point range). Cases with discrepancies of <50 points were recorded and reconciled on a two-headed microscope. Final H scores for each case were averaged by each pathologist. The expression scale of SHARPIN was graded by H score as follows: Low, H score 1–100; moderate, H score 101–200; and high, H score 201–300.

Statistical analysis

Statistical analysis was performed using SPSS 13.0 (SPSS, Inc., Chicago, IL, USA). Data were presented as the mean ± standard deviation. Differences in SHARPIN expression levels between normal skin and SCC, BCC and KA samples were analyzed using one-way analysis of variance and Tamhane's T2 post hoc test. The Broder grading system of SCC is commonly utilized to assess prognosis. It divides SCC into four categories based on histological grade. Grade I is composed of well-differentiated tumors, in which 75–100% of squamous cells are differentiated. Grade II is composed of moderately differentiated tumors in which 50–75% of squamous cells are differentiated. Grade III is composed of poorly differentiated tumors in which only 25–50% of cells are differentiated. Grade IV is an anaplastic tumor in which 0–25% of cells are differentiated (26). Main histologic variants of BCC include nodular type, adenoidal type, superficial type and sclerosing type (27). Associations between SHARPIN expression levels and aforementioned clinicopathological parameters were analyzed using the χ2 test for categorical variables. P<0.05 was considered to indicate a statistically significant difference.

Results

SHARPIN is aberrantly decreased in human NMSC

The SHARPIN protein was absent in the tumor nests and significantly decreased in precancerous lesions of SCC and BCC (Fig. 1) when compared to normal epithelium (Fig. 2). In addition, SHARPIN was moderately to highly expressed in KA samples (Fig. 3).

In BCC, SHARPIN expression was low in 63 cases (74.5%) and moderate in 22 cases (25.5%). In SCC, SHARPIN expression was low in 52 cases (68.1%) and moderate in 25 cases (31.9%). Furthermore, the difference in SHARPIN expression levels between BCC and normal skin, SCC and normal skin, and SCC and KA were all significant (P<0.05) (Fig. 4). However, no significant association was observed between SHARPIN expression and tumor grading of SCC. Demographics of all the patients and their H scores are summarized in Tables IIIVI.

Table III.

Demographics and H scores of patients with basal cell carcinoma.

Table III.

Demographics and H scores of patients with basal cell carcinoma.

IDSexAge, yLocationTypeH score
B01M74Noseadenoidal type10
B02F52Right earadenoidal type10
B03M72Trunksuperficial type10
B04F70Right handsuperficial type10
B05M68Right shouldersuperficial type10
B06M62Upper lipnodular type10
B07M63Noseadenoidal type10
B08M44Lower lipsclerosing type10
B09F82Lower lipsclerosing type10
B10F45Headadenoidal type10
B11F70Noseadenoidal type15
B12F63Right hipadenoidal type20
B13F49Right thighsclerosing type20
B14F73Left forearmsuperficial type20
B15M69Upper lipnodular type20
B16M69Noseadenoidal type20
B17M65Left handsclerosing type20
B18M67Noseadenoidal type20
B19F47Noseadenoidal type20
B20M65Necksuperficial type20
B21F73Headnodular type25
B22M22Nosenodular type30
B23M58Lower lipadenoidal type30
B24F58Nosesuperficial type30
B25F57Left cheekadenoidal type35
B26M59Upper lipadenoidal type35
B27F68Right tempussuperficial type40
B28F80Upper lippigmented type40
B29F95Backnodular type40
B30M78Backnodular type40
B31F65Left tempusnodular type40
B32F76Backpigmented type40
B33M64Left legsuperficial type40
B34M43Headsclerosing type45
B35F41Left foreheadpigmented type50
B36F68Chestpigmented type50
B37F83Right handnodular type50
B38M89Nosesuperficial type50
B39F75Left handpigmented type50
B40F75Lower lipsuperficial type55
B41M63Nosesclerosing type55
B42M45Upper lipnodular type60
B43F64Left thighadenoidal type60
B44M50Backnodular type60
B45M59Lower lipnodular type60
B46M46Upper lipadenoidal type70
B47M81Headnodular type70
B48F60Lower lipnodular type70
B49M58Left thighadenoidal type70
B50F43Anusnodular type70
B51F78Left thighnodular type75
B52F61Nosenodular type80
B53M56Nosenodular type80
B54M53Left thighnodular type80
B55F45Right cheekpigmented type80
B56F29Left checksuperficial type80
B57F44Left thighadenoidal type80
B58F58Left thighnodular type80
B59F73Backnodular type85
B60M58Nosesuperficial type90
B61F73Headnodular type90
B62M73Left cheeksclerosing type90
B63F57Nosenodular type90
B64F34Lower lipsuperficial type100
B65F41Nosenodular type100
B66F34Lower lipsuperficial type100
B67M73Lower lipnodular type100
B68M82Nosenodular type100
B69M63Backnodular type100
B70F38Upper lipsuperficial type105
B71F53Right tempusnodular type105
B72F64Backnodular type110
B73F53Lower lipnodular type120
B74F63Backnodular type120
B75M35Upper lipsuperficial type120
B76F44Nosesuperficial type130
B77F60Right tempusnodular type135
B78F70Nosesuperficial type140
B79M37Nosenodular type140
B80F36Nosenodular type140
B81F64Right tempusadenoidal type150
B82M56Lower lipsclerosing type150
B83M73Backsuperficial type160
B84F47Right handnodular type160
B85M58Lower lipnodular type190

[i] F, female; M, male.

Table VI.

Demographics and H scores of negative control patients.

Table VI.

Demographics and H scores of negative control patients.

IDSexAge, yH score
N01M30220
N02F22260
N03F57260
N04F48275
N05M32280
N06F50240
N07M40280
N08M18245
N09M39245
N10F28250
N11F69270
N12M55250

[i] F, female; M, male.

SHARPIN mutation analysis

A total of 8 exons and exon-intron adjacent sequences of SHARPIN were analyzed using DNA extracts from FFPE blocks of BCC, SCC and KA samples and healthy skin specimens, and DNA extracts from peripheral blood samples of 100 normal controls. Complete descriptions of the mutations detected in BCC and SCC are presented in Table VII. Total mutation rates were 21.8% in BCC and 17.0% in SCC samples. The C>T substitutions were 5.5% in BCC and 6.4% in SCC. Additionally, no mutations of SHARPIN were detected in DNA extracts from one benign skin tumor, 12 healthy skin tissues and blood samples from 100 normal individuals.

Table VII.

Distribution of Src homology 3 and multiple ankyrin repeat domains protein-associated RH domain-interacting protein gene mutations in patients with BCC and SCC.

Table VII.

Distribution of Src homology 3 and multiple ankyrin repeat domains protein-associated RH domain-interacting protein gene mutations in patients with BCC and SCC.

Tumor typeExonMutationModified proteinFrequencyDomain
BCCE1c.10 C>Tp.Pro4Leu1/55
c.68 C>Tp.Ala23Val1/55
c.146 A>Gp.Asp49Gly1/55
E2c.329 T>Cp.Gln110Arg1/55PH
E5c.733 C>Ap.His245Thr1/55UBL
E7c.937 C>Tp.Pro313Ser1/55
c.944 A>Gp.His315Arg1/55
c.992 T>Cp.Leu332Ser3/55
E8c.1109 T>Cp.Met370Thr1/55NZF
c.1137 G>Ap.Trp379Gln1/55
SCCE1c.53 C>Ap.Ala18Asp1/47
E2c.214 T>Cp.Trp72Arg1/47PH
E3c.421 C>Tp.Pro141Ser1/47
c.466 C>Tp.Pro156Ser1/47
c.469 C>Tp.Pro157Ser1/47
c.478 G>Ap.Ala160Thr1/47
E5c.709 T>Cp.Ser237Pro1/47
E8c.1007 G>Tp.Gly336Val1/47

[i] BCC, basal cell carcinoma; SCC, squamous cell carcinoma; PH, pleckstrin homology domain; UBL, ubiquitin-like; NZF, Npl4 zinc finger.

Discussion

The present study evaluated the expression of SHARPIN protein and analyzed the sequences of SHARPIN in NMSC. To the best of our knowledge, this is the first study to comprehensively investigate the expression and mutations of SHARPIN in a large series of patients with NMSC.

The essential contribution of SHARPIN to the activation of NF-κB supports the possibility that SHARPIN promotes tumorigenesis, as NF-κB signaling possesses well-demonstrated tumorigenic properties (12). This is supported by the SHARPIN-mediated suppression of apoptosis in the keratinocytes and hepatocytes of cpdm mice (18). Additionally, SHARPIN promotes the migration of Chinese hamster ovary cells in vitro and lymphocytes in vivo, and increases the lung metastasis of osteosarcoma in vivo in immunocompromised mice (19). In addition, the upregulation of SHARPIN has been observed in different types of internal solid cancer, including ovarian cancer, renal cell carcinoma, and cervical and prostate cancer (20,21). Furthermore, SHARPIN induces PTEN polyubiquitination independently of the K48 linkage. This process requires the UBL domain, which mediates SHARPIN's association with PTEN and its ability to bind ubiquitin via the NZF motif (28). Rantala et al (16) demonstrated that SHARPIN inactivates integrins in a number of different cell types and affects integrin-dependent cellular functions. Bii et al (22) identified SHARPIN as a metastasis gene in breast cancer using a replication-incompetent gammaretroviral vector, suggesting the potential of SHARPIN as a biomarker for stratifying patients with breast cancer. Additionally, Haris et al (23) identified that SHARPIN was significantly upregulated in U87 glioblastoma cells upon treatment with Aloe-emodin. Collectively, substantial evidence has demonstrated the role of SHARPIN in promoting tumorigenesis. Despite these data, a PubMed search did not identify any studies examining the expression of SHARPIN in NMSC. Therefore, the present study explored the expression of SHARPIN in three types of skin tumors, including the malignant forms BCC and SCC.

Firstly, the expression of SHARPIN was detected via immunohistochemistry. Contrary to the results of examination of internal solid tumors (17), SHARPIN expression was downregulated or absent in the majority of NMSC samples compared with normal skin tissues and KA. KA is commonly diagnosed clinically as it rapidly appears and develops as a raised lesion; however, as a non-pigmented lesion with a central keratin plug, SCC may also exhibit the same appearance. Furthermore, cases of KA with SCC arising from the base have been identified (29). Differential diagnosis between KA and SCC is challenging due to their similarities and the lack of reliable diagnostic criteria to distinguish them. Therefore, whether KA is a separate benign entity, or a variant of SCC, is controversial. At present, no biomarkers exist to distinguish SCC from KA, and KA lesions are commonly treated in the same way as SCC. However, SCC has a poorer prognosis than KA and is treated more aggressively; therefore, distinguishing between these two malignancies would be advantageous in order to implement the appropriate treatment, thereby decreasing unnecessary surgeries, the burden on the healthcare system and, importantly, the anxiety of the patients (30). Based on the results of the present study, that SHARPIN is absent or exhibits low expression in SCC but a high expression in KA, we hypothesize that SHARPIN may allow early differentiation and in situ treatment of SCC and KA to avoid metastasis and tissue destruction of SCC and the overtreatment of KA.

At present, the mechanism of how the downregulation of SHARPIN promotes skin tumorigenesis remains to be elucidated. Ikeda et al (11) identified that the absence of SHARPIN in cpdm mice caused dysregulation of NF-κB and increased apoptosis and necrosis of mouse embryonic fibroblasts. The data from the present study suggested that the downregulation of SHARPIN in NMSC may impair the function of LUBAC, and subsequently, the activation of NF-κB. In the majority of tumors, the aberrant activation of NF-κB signaling stimulates tumor cell proliferation, invasion and metastasis (31). Counterintuitively, Hogerlinden et al (32) demonstrated that selective inhibition of Rel/NF-κB signaling in the skin leads to disturbed epidermal homeostasis and hair follicle development, increased frequency of apoptotic keratinocytes and spontaneous development of SCC. Notably, a number of data have challenged the view that apoptotic signaling solely serves to inhibit cancer, arguing instead that apoptosis is responsible for various effects that may be tumor-promoting (3336). Apoptotic cell death is a cell-autonomous event, but its effects are not; dying cells affect their surrounding environments in various ways, which may include stimulating the proliferation of neighboring cells, affecting intra-tumoral cell competition and exerting paracrine effects on tumor microenvironments. Various data support the hypothesis that apoptosis may promote tumorigenesis through the recruitment and activation of phagocytic macrophages at the tumor sites (37). Taken together, we hypothesized that decreased SHARPIN expression may promote NMSC through the impaired activation of NF-κB and increased apoptosis and necrosis of epidermal cells, which may disrupt the homeostasis of the epidermis and lead to tumorigenesis.

Traditional Sanger sequencing has been the gold standard for identifying mutations for a number of years due to its low false-positive rate and high specificity (21). Therefore, in the present study, DNA was extracted from NMSC FFPE blocks and mutations in the exons of SHARPIN were detected. The results indicated that high proportions of BCC and SCC contained mutations of the SHARPIN gene. Mutations in SHARPIN exons were identified in 21.8% of BCC and 17.0% of SCC in the present study. The proportions of C>T substitutions were 5.5% in BCC and 6.4% in SCC samples, which were identified as characteristic of mutations associated with exposure to UV exposure (38). In addition, the mutations were not only located in the UBL domain of SHARPIN but also in the PH and NZF domains, thereby potentially affecting other functions of SHARPIN besides the formation of LUBAC. Furthermore, SHARPIN has been indicated to inactivate integrins in a number of cell types and affect integrin-dependent cellular functions independent of LUBAC (16). Approximately one-half of the cellular SHARPIN is not associated with the LUBAC complex (28). Therefore, the present study concluded that SHARPIN is a multifunctional molecule and may promote the pathogenesis of NMSC through different mechanisms.

Overall, the present study contributes to a growing body of evidence supporting the importance of SHARPIN in NMSC. The results suggest an association between NMSC and low to absent SHARPIN expression and SHARPIN mutations.

It was identified that SHARPIN protein expression was absent in cancer nests and significantly decreased in precancerous lesions of SCC and BCC, but was high in normal skin or in KA. The total mutation rates of SHARPIN were 21.8% in BCC and 17.0% in SCC. These data indicated that SHARPIN may serve a tumor-suppressing role and act as a promising diagnostic biomarker in NMSC.

Acknowledgements

Not applicable.

Funding

The present study was supported by the National Natural Science Foundation of China (grant no., 81371724) to Dr Yanhua Liang.

Availability of data and material

The datasets used and/or analyzed during the present study are available from the corresponding authors on reasonable request.

Authors' contributions

YL designed and supervised the study. YZ performed the histological examination of the samples and prepared the draft. YY performed the DNA extraction, polymerase chain reaction and sequencing. JW conducted data analysis and interpreted the data.

Ethics approval and consent to participate

Blood specimens of 100 normal individuals and skin tissues from 12 healthy volunteers who received cosmetic surgeries, and formalin-fixed paraffin-embedded (FFPE) samples from the Department of Dermatology of Shenzhen Hospital in Southern Medical University (Shenzhen, China), were obtained according to a protocol approved by the Southern Medical University Shenzhen Hospital Subject Review Board (2016–016). All samples were fixed for 24 h in 10% formalin solution at room temperature. The thickness of the sections was 4 µm. Informed consent was obtained from all participants.

Patient consent for publication

Consent was gained from the participants for the publication of their data.

Competing interests

The authors declare that they have no competing interests.

References

1 

Rodust PM, Stockfleth E, Ulrich C, Leverkus M and Eberle J: UV-induced squamous cell carcinoma-a role for antiapoptotic signalling pathways. Br J Dermatol. 161 (Suppl 3):S107–S115. 2009. View Article : Google Scholar

2 

Molho-Pessach V and Lotem M: Ultraviolet radiation and cutaneous carcinogenesis. Curr Probl Dermatol. 35:14–27. 2007. View Article : Google Scholar : PubMed/NCBI

3 

de Villiers EM, Ruhland A and Sekarić P: Human papillomaviruses in non-melanoma skin cancer. Semin Cancer Biol. 9:413–422. 1999. View Article : Google Scholar : PubMed/NCBI

4 

McNaughton SA, Marks GC and Green AC: Role of dietary factors in the development of basal cell cancer and squamous cell cancer of the skin. Cancer Epidemiol Biomarkers Prev. 14:1596–1607. 2005. View Article : Google Scholar : PubMed/NCBI

5 

Ramsay HM, Fryer AA, Reece S, Smith AG and Harden PN: Clinical risk factors associated with nonmelanoma skin cancer in renal transplant recipients. Am J Kidney Dis. 36:167–176. 2000. View Article : Google Scholar : PubMed/NCBI

6 

Giglia-Mari G and Sarasin A: TP53 mutations in human skin cancers. Hum Mutat. 21:217–228. 2003. View Article : Google Scholar : PubMed/NCBI

7 

Madan V, Lear JT and Szeimies RM: Non-melanoma skin cancer. Lancet. 375:673–685. 2010. View Article : Google Scholar : PubMed/NCBI

8 

Eisemann N, Jansen L, Castro FA, Chen T, Eberle A, Nennecke A, Zeissig SR, Brenner H and Katalinic A; GEKID Cancer Survival Working Group, : Survival with nonmelanoma skin cancer in Germany. Br J Dermatol. 174:778–785. 2016. View Article : Google Scholar : PubMed/NCBI

9 

Stieglitz B, Haire LF, Dikic I and Rittinger K: Structural analysis of SHARPIN, a subunit of a large multi-protein E3 ubiquitin ligase, reveals a novel dimerization function for the pleckstrin homology superfold. J Biol Chem. 287:20823–20829. 2012. View Article : Google Scholar : PubMed/NCBI

10 

Seymour RE, Hasham MG, Cox GA, Shultz LD, Hogenesch H, Roopenian DC and Sundberg JP: Spontaneous mutations in the mouse Sharpin gene result in multiorgan inflammation, immune system dysregulation and dermatitis. Genes Immun. 8:416–421. 2007. View Article : Google Scholar : PubMed/NCBI

11 

Ikeda F, Deribe YL, Skånland SS, Stieglitz B, Grabbe C, Franz-Wachtel M, van Wijk SJ, Goswami P, Nagy V, Terzic J, et al: SHARPIN forms a linear ubiquitin ligase complex regulating NF-κB activity and apoptosis. Nature. 471:637–641. 2011. View Article : Google Scholar : PubMed/NCBI

12 

Tokunaga F, Nakagawa T, Nakahara M, Saeki Y, Taniguchi M, Sakata S, Tanaka K, Nakano H and Iwai K: SHARPIN is a component of the NF-κB-activating linear ubiquitin chain assembly complex. Nature. 471:633–636. 2011. View Article : Google Scholar : PubMed/NCBI

13 

Haas AL: Linear polyubiquitylation: The missing link in NF-kappaB signalling. Nat Cell Biol. 11:116–118. 2009. View Article : Google Scholar : PubMed/NCBI

14 

Tokunaga F, Sakata S, Saeki Y, Satomi Y, Kirisako T, Kamei K, Nakagawa T, Kato M, Murata S, Yamaoka S, et al: Involvement of linear polyubiquitylation of NEMO in NF-kappaB activation. Nat Cell Biol. 11:123–132. 2009. View Article : Google Scholar : PubMed/NCBI

15 

Wang Z, Potter CS, Sundberg JP and Hogenesch H: SHARPIN is a key regulator of immune and inflammatory responses. J Cell Mol Med. 16:2271–2279. 2012. View Article : Google Scholar : PubMed/NCBI

16 

Rantala JK, Pouwels J, Pellinen T, Veltel S, Laasola P, Mattila E, Potter CS, Duffy T, Sundberg JP, Kallioniemi O, et al: SHARPIN is an endogenous inhibitor of β1-integrin activation. Nat Cell Biol. 13:1315–1324. 2011. View Article : Google Scholar : PubMed/NCBI

17 

Jung J, Kim JM, Park B, Cheon Y, Lee B, Choo SH, Koh SS and Lee S: Newly identified tumor-associated role of human Sharpin. Mol Cell Biochem. 340:161–167. 2010. View Article : Google Scholar : PubMed/NCBI

18 

Sieber S, Lange N, Kollmorgen G, Erhardt A, Quaas A, Gontarewicz A, Sass G, Tiegs G and Kreienkamp HJ: Sharpin contributes to TNFα dependent NFκB activation and anti-apoptotic signalling in hepatocytes. PLoS One. 7:e299932012. View Article : Google Scholar : PubMed/NCBI

19 

De Melo J and Tang D: Elevation of SIPL1 (SHARPIN) increases breast cancer risk. PLoS One. 10:e01275462015. View Article : Google Scholar : PubMed/NCBI

20 

Zhang Y, Huang H, Zhou H, Du T, Zeng L, Cao Y, Chen J, Lai Y, Li J, Wang G, et al: Activation of nuclear factor κB pathway and downstream targets survivin and livin by SHARPIN contributes to the progression and metastasis of prostate cancer. Cancer. 120:3208–3218. 2014. View Article : Google Scholar : PubMed/NCBI

21 

Li Q, Li M, Ma L, Li W, Wu X, Richards J, Fu G, Xu W, Bythwood T, Li X, et al: A method to evaluate genome-wide methylation in archival formalin-fixed, paraffin-embedded ovarian epithelial cells. PLoS One. 9:e1044812014. View Article : Google Scholar : PubMed/NCBI

22 

Bii VM, Rae DT and Trobridge GD: A novel gammaretroviral shuttle vector insertional mutagenesis screen identifies SHARPIN as a breast cancer metastasis gene and prognostic biomarker. Oncotarget. 6:39507–39520. 2015. View Article : Google Scholar : PubMed/NCBI

23 

Haris K, Ismail S, Idris Z, Abdullah JM and Yusoff AA: Expression profile of genes modulated by Aloe emodin in human U87 glioblastoma cells. Asian Pac J Cancer Prev. 15:4499–4505. 2014. View Article : Google Scholar : PubMed/NCBI

24 

Ghaheri M, Kahrizi D, Yari K, Babaie A, Suthar RS and Kazemi E: A comparative evaluation of four DNA extraction protocols from whole blood sample. Cell Mol Biol. 62:120–124. 2016.PubMed/NCBI

25 

Bollag G, Hirth P, Tsai J, Zhang J, Ibrahim PN, Cho H, Spevak W, Zhang C, Zhang Y, Habets G, et al: Clinical efficacy of a RAF inhibitor needs broad target blockade in BRAF-mutant melanoma. Nature. 467:596–599. 2010. View Article : Google Scholar : PubMed/NCBI

26 

Kallini JR, Hamed N and Khachemoune A: Squamous cell carcinoma of the skin: Epidemiology, classification, management, and novel trends. Int J Dermatol. 54:130–140. 2015. View Article : Google Scholar : PubMed/NCBI

27 

Paolino G, Donati M, Didona D, Mercuri SR and Cantisani C: Histology of non-melanoma skin cancers: An update. Biomedicines. 5(pii): E712017. View Article : Google Scholar : PubMed/NCBI

28 

De Melo J, Lin X, He L, Wei F, Major P and Tang D: SIPL1-facilitated PTEN ubiquitination contributes to its association with PTEN. Cell Signal. 26:2749–2756. 2014. View Article : Google Scholar : PubMed/NCBI

29 

Weedon DD, Malo J, Brooks D and Williamson R: Squamous cell carcinoma arising in keratoacanthoma: A neglected phenomenon in the elderly. Am J Dermatopathol. 32:423–426. 2010. View Article : Google Scholar : PubMed/NCBI

30 

Vasiljević N, Andersson K, Bjelkenkrantz K, Kjellström C, Månsson H, Nilsson E, Landberg G, Dillner J and Forslund O: The Bcl-xL inhibitor of apoptosis is preferentially expressed in cutaneous squamous cell carcinoma compared with that in keratoacanthoma. Int J Cancer. 124:2361–2366. 2009. View Article : Google Scholar : PubMed/NCBI

31 

Karin M and Greten FR: NF-kappaB: Linking inflammation and immunity to cancer development and progression. Nat Rev Immunol. 5:749–759. 2005. View Article : Google Scholar : PubMed/NCBI

32 

van Hogerlinden M, Rozell BL, Ahrlund-Richter L and Toftgård R: Squamous cell carcinomas and increased apoptosis in skin with inhibited Rel/nuclear factor-kappaB signaling. Cancer Res. 59:3299–3303. 1999.PubMed/NCBI

33 

Michalak EM, Vandenberg CJ, Delbridge AR, Wu L, Scott CL, Adams JM and Strasser A: Apoptosis-promoted tumorigenesis: Gamma-irradiation-induced thymic lymphomagenesis requires Puma-driven leukocyte death. Genes Dev. 24:1608–1613. 2010. View Article : Google Scholar : PubMed/NCBI

34 

Li F, Huang Q, Chen J, Peng Y, Roop DR, Bedford JS and Li CY: Apoptotic cells activate the ‘phoenix rising’ pathway to promote wound healing and tissue regeneration. Sci Signal. 3:ra132010. View Article : Google Scholar : PubMed/NCBI

35 

Huang Q, Li F, Liu X, Li W, Shi W, Liu FF, O'Sullivan B, He Z, Peng Y, Tan AC, et al: Caspase 3-mediated stimulation of tumor cell repopulation during cancer radiotherapy. Nat Med. 17:860–866. 2011. View Article : Google Scholar : PubMed/NCBI

36 

Qiu W, Wang X, Leibowitz B, Yang W, Zhang L and Yu J: PUMA-mediated apoptosis drives chemical hepatocarcinogenesis in mice. Hepatology. 54:1249–1258. 2011. View Article : Google Scholar : PubMed/NCBI

37 

Ichim G and Tait SW: A fate worse than death: Apoptosis as an oncogenic process. Nat Rev Cancer. 16:539–548. 2016. View Article : Google Scholar : PubMed/NCBI

38 

Scott GA, Laughlin TS and Rothberg PG: Mutations of the TERT promoter are common in basal cell carcinoma and squamous cell carcinoma. Mod Pathol. 27:516–523. 2014. View Article : Google Scholar : PubMed/NCBI

Related Articles

Journal Cover

April-2019
Volume 17 Issue 4

Print ISSN: 1792-0981
Online ISSN:1792-1015

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Zheng Y, Yang Y, Wang J and Liang Y: Aberrant expression and high‑frequency mutations of SHARPIN in nonmelanoma skin cancer. Exp Ther Med 17: 2746-2756, 2019
APA
Zheng, Y., Yang, Y., Wang, J., & Liang, Y. (2019). Aberrant expression and high‑frequency mutations of SHARPIN in nonmelanoma skin cancer. Experimental and Therapeutic Medicine, 17, 2746-2756. https://doi.org/10.3892/etm.2019.7261
MLA
Zheng, Y., Yang, Y., Wang, J., Liang, Y."Aberrant expression and high‑frequency mutations of SHARPIN in nonmelanoma skin cancer". Experimental and Therapeutic Medicine 17.4 (2019): 2746-2756.
Chicago
Zheng, Y., Yang, Y., Wang, J., Liang, Y."Aberrant expression and high‑frequency mutations of SHARPIN in nonmelanoma skin cancer". Experimental and Therapeutic Medicine 17, no. 4 (2019): 2746-2756. https://doi.org/10.3892/etm.2019.7261