toxicol. sci. 2003 chipman 146 53

Upload: kash-evangelina

Post on 03-Jun-2018

217 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/12/2019 Toxicol. Sci. 2003 Chipman 146 53

    1/8

    FORUM

    Disruption of Gap Junctions in Toxicity and Carcinogenicity

    J. Kevin Chipman,1

    Angela Mally, and Gareth O. Edwards

    School of Biosciences, The University of Birmingham, Edgbaston, Birmingham B15 2TT, United Kingdom

    Received October 11, 2002; accepted November 16, 2002

    Although the specific role of connexin-mediated gap junctional

    intercellular communication in the control of cell homeostasis,

    proliferation, and death are still not clear, several lines of evidence

    support these roles. The disturbance of this communication,

    through multiple mechanisms, may in the short term be a protec-tive mechanism to limit the spread of toxicity in a tissue following

    chemical or radiation damage. However, sustained downregula-

    tion confers a loss of tumor-suppressive action. Consequently,

    connexin dysfunction has been associated with both the action of

    many carcinogens and being a feature of cancerper se. Connexins

    offer not only a target for cancer chemoprevention but also for

    exploitation in chemotherapy through the bystander effect.

    Key Words: connexin, gap junction, carcinogenesis, bystander,

    intercellular communication.

    Structure and Function of Gap Junctions

    Gap junctions are intercellular membrane channels that al-low the direct exchange of small molecules (1.2 kD) between

    adjacent cells. They are comprised of two hemichannels (con-

    nexons), which are in turn formed by the oligomerization of six

    protein subunits, termed connexins (Cxs). To date, at least 21

    members of this multigene family have been discovered and

    cloned (Willecke et al., 2002). Cxs consist of four highly

    conserved membrane-spanning domains, two extracellular

    loops, a cytoplasmic loop, and cytoplasmic N- and C-terminal

    regions. While the extracellular loops are thought to be re-

    quired for connexon docking, the cytoplasmic domains appear

    to play a role in channel gating and intracellular signaling, and

    they vary between members of the Cx family. Small ions such

    as water, Ca2

    , cAMP, and inositol triphosphate can passthrough the gap junctions from one cell to another, while larger

    molecules such as proteins, complex lipids, and polysaccha-

    rides are retained within the cell (Saez et al., 1989). This

    process of exchange of molecules between neighboring cells

    via gap junctions is termed gap junction intercellular commu-

    nication (GJIC). A major physiological role of GJIC is the

    maintenance of tissue homeostasis and proliferation, the regu-

    lation of embryonic development and differentiation, and elec-

    tric coupling of electrically excitable cells such as cardiomy-

    ocytes (Loewenstein, 1981). However, more recently, evidence

    for an intracellular signaling role of connexins in the control ofcell proliferation has also emerged (see below).

    Mechanisms of Regulation of GJIC

    Gap junction-mediated intercellular communication can be

    regulated at several levels, including transcription, mRNA

    processing, protein synthesis, and post-translational modifica-

    tion, assembly, trafficking, connexon docking, and gating of

    the gap junction channel (see Fig. 1).

    Several promoter response elements and transcription fac-

    tors that are involved in the transcriptional regulation of con-

    nexins have been identified. For example, the promoter of the

    rat cx32 gene contains several GC-boxes, and Sp1, HNF-1,

    YY-1, and NF-1 binding sites have been recognized (Bai et al.,

    1995; Piechocki et al, 2000). In the human, rat and, mouse

    cx43 promoters, several sites, including AP1, cAMP and es-

    trogen response elements have been identified (Chen et al.,

    1995a; Yu et al., 1994). The expression of cx genes can be

    controlled in a tissue-specific manner. For example, the liver

    cell-specific expression of cx32 has been shown to be depen-

    dent (at least in part) on expression of HNF-1 (Piechocki et al.,

    2000). Additional promoters have been identified that control

    the transcription of cx32 in neuronal tissue (Neuhaus et al.,

    1995). The methylation of CpG islands of promoter regions

    can lead to gene silencing and methylation of the Cx43 pro-

    moter has been reported in MH1C1 rat hepatoma cells, whichexpress Cx32 but not Cx43, while the cx32 promoter was

    methylated in WB-F344 rat liver epithelial cells, which express

    Cx43 but not Cx32 (Piechocki et al., 1999).

    It appears that connexin mRNA stability also plays an im-

    portant role in controlling GJIC. Decreased Cx32 mRNA half-

    life has been observed in regenerating liver (Kren et al., 1993),

    while sequences located within the 3 untranslated region of

    the Cx43 mRNA appear to stabilize this transcript (Lefebvre et

    al.,1995).

    1 To whom correspondence should be addressed. Fax: 44 121 414 5925.

    E-mail: [email protected].

    TOXICOLOGICAL SCIENCES71, 146153 (2003)

    Copyright 2003 by the Society of Toxicology

    146

  • 8/12/2019 Toxicol. Sci. 2003 Chipman 146 53

    2/8

    The activity of GJ is also regulated by post-translational

    modification of connexin proteins. Regulation can be by volt-

    age, calcium and various kinases that act on the intracellular

    domains (Fig. 1; Holder et al., 1993). Phosphorylation of the

    C-terminal domain of the connexins (with the exception of

    Cx26) influences Cx trafficking, channel gating and turnover

    (Saezet al., 1998). Inhibition of communication through Cx43

    hyperphosphorylation has been brought about both chemically

    (phorbol ester, via MAP kinase and possibly protein kinase C)

    and by epidermal growth factor through the action of MAP

    kinase (Rivedal and Opsahl, 2001). Hepatic Cx32 appears also

    to be inactivated by PKC activation (Elcock et al., 2000).

    Furthermore a number of oncogene products mediate inhibitionof GJIC, such as through the tyrosine phoshorylation of Cx43

    by v-src (reviewed by Yamasaki et al.,1999). However, there

    are multiple phosphorylation sites on different Cx molecules

    and the balance of the phosphorylation status determines func-

    tion. It is predictable, therefore, that phosphorylation is medi-

    ated by various signal transduction pathways.

    Another important mechanism involves the trafficking of

    connexins from intracellular pools to the plasma membrane.

    Following protein synthesis a number of the connexin polypep-

    tides have been shown to be inserted into the membrane of the

    ER and subsequently the Golgi where they oligomerize to form

    hemichannels. Little is known about what initiates and medi-

    ates the transport of connexons to the plasma membrane and

    the formation of the complete gap junction channel, but cell

    cell adhesion appears to be especially important. There is a

    requirement for E-cadherin expression for Cx43 transport to

    the cell membrane and for communication to be recovered in a

    mouse skin papilloma cell line (Hernandez-Blazquez et al.,

    2001). It is also known that Cxs can associate with other

    proteins in the membrane, specifically ZO-1 with Cx43 and

    Cx45 (Defamie et al.,2001; Kausalya et al.,2001). Clustering

    of gap junction channels into plaques is required for functionalcell coupling (Bukauskas et al., 2000). Finally, several mech-

    anisms have been implicated in GJ degradation, including

    lysosomal and proteasomal pathways.

    It is therefore not surprising to find that there are multiple

    ways by which GJIC can be regulated and disrupted (Saez et

    al.,1993). Aliphatic alcohols (octanol, heptanol) and anesthet-

    ics such as halothane impair gap junction activity by interfering

    with membrane fluidity that leads to contraction of intercell

    channels. Other compounds have been reported to physically

    FIG. 1. Summary of the multiple regulatory points that control GJIC. Specificity of regulation and function are dependent on the connexin forms.

    147FORUM

  • 8/12/2019 Toxicol. Sci. 2003 Chipman 146 53

    3/8

    block GJs. Chemical gating is regulated by pH, calcium ions

    and free radicals (Saez et al., 1993). Thus, toxicant-induced

    changes in cellular calcium levels and redox state can lead to

    inhibition of GJIC at least in part through mechanisms that

    involve disturbance of phophorylation status (Trosko et al.,

    1998).

    GJIC and Control of Cell Proliferation and Apoptosis:

    Multiple Signaling Roles

    GJIC has long been considered to play an important role in

    the regulation of cell growth and cell death. For example, in the

    course of liver regeneration following partial hepatectomy, GJ

    expression is temporarily decreased until liver mass and lobu-

    lar conformation are restored. Similar effects on GJIC are seen

    during acute liver injury and tissue restoration after treatment

    with toxic chemicals such as thioacetamide (Kojima et al.,

    1994). Studies on liver regeneration after partial hepatectomy,

    where synchronous initiation of DNA synthesis was preceded

    by a decrease in the expression of Cx proteins, suggest thatdownregulation of GJs may be involved as a trigger for cell

    proliferation (Dermietzelet al.,1987). However, Temme et al.

    (2000) reported that Cx32-deficient mice showed normal re-

    generative response following two-thirds hepatectomy, although

    the extent of initiation and termination of DNA synthesis was

    slightly altered. Furthermore, hepatocytes completely devoid

    of gap junctional communication are able to enter the cell cycle

    upon mitogenic stimulation (Fladmarket al., 1997). Neveu et

    al. (1990) demonstrated that a decrease of Cx32 in livers of

    phenobarbitone-treated rats was not associated with increased

    hepatocyte proliferation. Similarly, we have recently shown

    that downregulation of Cxs by a number of nongenotoxic

    carcinogens does not necessarily correlate with induction of

    cell proliferation (Mally and Chipman, 2002). Although a

    certain stage of mitosis has been associated with decreased

    levels of Cxs (Dermietzel et al., 1987), loss of GJIC per se

    does not appear to provide a signal for cells to divide but rather

    permits cell-cycle progression in the presence of mitotic stim-

    uli (Trosko and Ruch 1998). It has been suggested that down-

    regulation of GJs during mitosis may serve to maintain sepa-

    rate cellular pools of signaling molecules (Fladmark et al.,

    1997). This would ensure that during critical stages of cell

    division, proliferating cells remain isolated from potentially

    damaging signals mediated by neighboring cells.

    Just how gap junctions regulate cell growth is poorly under-stood. Forced expression of Cx43 and Cx32 in neoplastic lung

    and liver cells restored G1 growth control and was associated

    with increased p27 (kip-1) and decreased cyclin D1 expression

    (Koffler et al., 2000). Similarly, overexpression of Cx43 in

    osteosarcoma cells inhibited cell cycle transition from G1 to S

    phase by increasing the expression of p27 (Zhanget al., 2001).

    Chen et al. (1995b) reported alterations in the expression of

    genes involved in cell cycle control after transfection of Cx43

    into a transformed kidney cell line. However, it remains to be

    established whether the observed decrease in cyclin A, D1 and

    D2 and the cyclin dependent kinases CDK5 and CDK6 were a

    direct effect of Cx43 expression or a result of the accompanied

    density-dependent inhibition of proliferation and prolongation

    of the G1 and S phases of the cell cycle.

    It is becoming increasingly apparent that gap junctions play

    an important role in intracellular signaling in relation to pro-liferation in addition to junctional functionality (Duflot-Dancer

    et al., 1997). Moorby and Patel (2001) reported opposing

    effects on cell growth in three clones transfected with different

    Cx43 mutants when grown under noncoupling conditions. In

    addition, they were able to show that the cytoplasmic carboxyl

    domain of Cx43, which includes various phosphorylation sites

    and has no channel activity, is as effective as wild-type Cx43

    in blocking cell growth. These studies suggest that Cxs may

    regulate cell growth independent of intercell communication.

    As mitosis and apoptosis share some common regulatory

    features (Kasten and Giordano, 1998), it is not surprising that

    GJIC appears to play a role in both processes. Wilson et al.

    (2000) reported relatively high levels of GJIC during earlystages of apoptosis and mitosis, which then decreased as each

    process progressed. During both processes, activation of signal

    transduction pathways involving Cdc2/cyclin B kinase is an

    important event (King and Cidlowski, 1995; Pandey and

    Wang, 1995) and the phosphorylation of Cx43 in vitro is

    dependent on Cdc2 kinase (Lampe et al.,1998). However, the

    question remains whether phosphorylation of Cx43 influences

    apoptosis and mitosis or whether initiation of these pathways

    results in Cx phosphorylation and inhibition of GJIC.

    Since several tumor promoters and oncogenes are known to

    inhibit both GJIC and apoptosis, while several antitumor pro-

    moters enhance GJIC and apoptosis, Trosko and Goodman

    (1994) hypothesized that a death signal may be mediated

    through gap junctions. The ability of nongenotoxic carcinogens

    such as the peroxisome proliferators and phorbol esters to

    disrupt GJIC and at the same time suppress apoptosis suggests

    a strong link between GJ and apoptosis in homeostatic control

    of solid tissue. However, although GJIC may be one of the

    mechanisms contributing to apoptosis signaling, there is little

    evidence to indicate that functional gap junctions are (gener-

    ally) required for apoptosis. We have recently investigated the

    effects of a number of nongenotoxic carcinogens on gap junc-

    tions in relation to apoptosis in target tissues in the rat. In this

    study, the peroxisome proliferator Wy-14,643 decreased both

    the number of hepatocyte gap junction plaques containingCx32 and the rate of hepatocyte apoptosis, while 2,3,7,8-

    tetrachlorodibenzo-p-dioxin (TCDD) appeared to downregu-

    late Cx32 without effects on apoptosis. Most importantly,

    treatment with methapyrilene led to a dramatic loss of connex-

    ins and increased rate of apoptosis (Mally and Chipman, 2002).

    Similarly, propylthiouracil and low-iodine diet decreased GJIC

    and at the same time enhanced apoptosis in rat thyroid (Kolaja

    et al.,2000b). These data clearly show that apoptosis can occur

    in the absence of functional GJs, despite the fact that with

    148 FORUM

  • 8/12/2019 Toxicol. Sci. 2003 Chipman 146 53

    4/8

    certain other compounds a correlation between reduced GJ

    function, elevated mitogenesis, and reduced apoptosis exists

    (Kolaja et al., 2000a). It may be that functional GJIC is

    required for the initial signaling in apoptosis but that this

    communication is lost at a later stage. In summary, it appears

    that rather than suppressing growth or signaling a cell to

    undergo apoptosis, gap junctions may serve to maintain opti-mal rates of cell growth and cell death. Disruption of GJIC per

    se may not result in alterations of the rates of proliferation or

    apoptosis but require the presence of additional stimuli (Trosko

    and Ruch, 1998). Such a stimulus may include the activation of

    a cellular oncogene or inactivation of a tumor suppressor in

    initiated or preneoplastic cells (Watanabe and Tonosaki, 1995).

    Downregulation of GJIC as an Adaptive,

    Protective Response to Toxicity

    While GJIC can be important in tissue homeostasis and

    integrated tissue responses, it follows that GJs have the poten-

    tial to propagate toxic responses also through the transfer ofmolecules such as superoxide and calcium ions, if elevated in

    the process of cell toxicity (e.g., in cerebral ischemia; Lin et

    al.,1998). It is likely, therefore, that the rapid, temporary, and

    reversible closure of GJ during cellular stress may be an

    important protective means for preventing the spread of cellu-

    lar toxicity (Peracchia et al., 2000; Saez et al. 1989). Closure

    of gap junctions is thought to prevent the spread of apoptotic

    and inflammatory signals between cells, and furthermore, heart

    and brain infarction sizes are reduced when cells are uncoupled

    (reviewed in Brosnan et al., 2001). Studies have demonstrated

    that when cells are irradiated with 0.31.0 cGy doses of-par-

    ticles (Azzamet al., 2001; Zhou et al., 2000), stress responses

    (mutations, micronucleus formation, p53 phosphorylation,p21waf1 induction) are observed in more cells than were directly

    exposed. Upon incubation of cells with GJIC blockers such as

    lindane, the number of responding cells was significantly re-

    duced. However, when communicating cells were irradiated

    with soft x-rays at sublethal doses in the range of 1 to 5 Gy,

    GJIC was reduced in a dose-dependent manner (Edwardset al.,

    2002). Thus, depending on dose, the inhibition of GJIC can

    limit the spread of stress-inducing molecules through this very

    same system. This is of importance when considering the dose

    dependency of effects of exposure to ionizing radiation.

    Sustained Aberrant Localization or Downregulation

    of Cxs Provides a Tumor-Promoting Stimulus

    While deliberate, temporary closure of gap junctions serves

    to protect healthy cells by preventing the spread of toxic

    molecules and the loss of metabolically important substances,

    sustained inhibition of GJIC may contribute to carcinogenesis

    or conditions such as reproductive, neurological, and cardio-

    vascular dysfunction through loss of homeostatic control. Un-

    timely inhibition of GJIC during critical stages of development

    may result in embryonic toxicity or teratogenesis. Indeed,

    many chemicals known to be tumor promoters, teratogens, or

    neurotoxins modulate GJIC (Elmore et al., 1987; Rosenkranz

    et al., 1988; Trosko et al., 2002).

    There is substantial evidence to suggest that sustained down-

    regulation of GJIC provides a tumor-promoting stimulus. Nor-

    mal, contact-inhibited cells exhibit functional GJIC, while

    transformed cells and many tumors are characterized by re-duced expression of GJs involving aberrant localization or

    downregulation of Cxs and nonfunctional homologous or het-

    erologous GJIC (Loewenstein and Rose, 1992; Trosko and

    Ruch, 1998; Yamasaki, 1990). Activation of some cellular

    oncogenes inhibits GJIC and induces malignant transformation

    (Jou et al., 1995; reviewed in Yamasaki et al., 1999), and

    tumor-promoting conditions such as wounding and necrosis are

    associated with decreased GJIC. Moreover, compelling evi-

    dence that links loss of GJ function to carcinogenesis has come

    from the observation that many nongenotoxic carcinogens and

    tumor-promoting agents inhibit GJIC. A range of nongenotoxic

    carcinogens has tested positive for inhibitory effect on GJICin

    vitro. These include the pesticides DDT, DDE, lindane, hep-

    tachlor, and dieldrin, the peroxisome proliferators nafenopin,

    Wy-14,643, clofibrate, and di-2-ethylhexyl-phthalate, and sev-

    eral other classical tumor promoters in rodents, such as phe-

    nobarbitone, the phorbol ester 12-O-tetradecanoylphorbol 13-

    acetate, TCDD, and polychlorinated biphenyls (Swierenga and

    Yamasaki, 1992). Importantly, inhibition of GJIC by tumor-

    promoting agents has been demonstrated in several in vivoand

    ex vivo studies (Ito et al., 1998; Kolaja et al., 2000a; Kru-

    tovskikhet al., 1995; Mally and Chipman, 2002; Tateno et al.,

    1994). A very important aspect that has received insufficient

    attention is the need to distinguish between specific inhibition

    of GJIC and the inhibition that may result secondarily to, orcoincidentally with, cell death. In the latter case, such as is seen

    with carbon tetrachloride in rat liver (Cowles et al., 2000),

    compensatory cell proliferation is likely to be a more important

    tumor-promoting influence than the inhibition of gap junctions.

    However, in many cases, the window between the doses re-

    quired for GJ disruption and loss of cell viability is substantial

    both in vitro and in vivo, and the disruption has often been

    shown to be reversible (e.g., Elcock et al., 2000; Mally and

    Chipman, 2002; Rivedal and Opsahl, 2001; Swierenga and

    Yamasaki, 1992).

    It is thought that chronic disruption of GJIC may release

    potentially initiated cells from growth constraints imposed by

    normal neighboring cells, resulting in clonal expansion and

    ultimately tumor formation (Fig. 2). Following withdrawal of

    the tumor-promoting stimulus, GJ function is restored and is

    associated with reduced size of chemically induced preneoplas-

    tic foci (Neveu et al., 1990). Reduced expression of GJs

    following treatment with nongenotoxic carcinogens appears to

    be a target organ-specific effect (Mally and Chipman, 2002;

    Mesnil et al., 1988). Furthermore, the ability of some com-

    pounds to modulate GJIC correlates well with sex- and species-

    149FORUM

  • 8/12/2019 Toxicol. Sci. 2003 Chipman 146 53

    5/8

    specific tumor incidence (Klaunig et al., 1989; Plante et al.,

    2002).Many nongenotoxic carcinogens have been shown to cause

    gene hypermethylation, and this may be important in the re-

    duced expression of certain tumor-suppressive genes (Sug-

    imura and Ushijima, 2000). It is known that methylation is one

    important mechanism for the control of Cx expression (see

    above); and, in particular, the Cx26 gene promoter is found to

    be hypermethylated (notably at the SP-1 site) in certain breast

    cancer cell lines (Tan et al., 2002), and treatment with the

    demethylating agent 5-aza-2-deoxycytidine resulted in re-ex-

    pression of Cx26.

    Connexin Genes and Tumor Suppression

    Although cx gene mutations in tumors are rare, there is a

    wealth of evidence to suggest that Cxs function as tumor

    suppressors. As indicated above, the loss of GJIC is a common

    feature of tumors and appears to be important in the mecha-

    nism of action of many nongenotoxic carcinogens. Dominant

    negative mutants of cx32 and cx26 genes also caused loss of

    growth control, although this effect in some instances may be

    independent of GJIC per se (Duflot-Dancer et al., 1997).

    Numerous studies demonstrate that transfection of Cxs into

    communication deficient tumor cells can restore growth control

    in vitro and in vivo (Chen et al., 1995b; Eghbali et al., 1991;

    Mehta et al., 1991; Omori and Yamasaki 1999; Rae et al.,

    1998; Yano et al., 2001). Reduction of the malignant pheno-type in some of these studies was associated with reduced

    saturation density, increased doubling times, reduced growth

    capacity in soft agar, and delayed onset of tumor formation in

    nude mice. Similarly, treatment of BALB/c 3T3 cells with

    Cx43 antisense oligonucleotides resulted in inhibition of GJIC

    and increased saturation density (Ruch et al.,1995). Dominant

    negative inhibition of GJIC by mutant Cx43 enhanced tumor-

    igenicity of a rat bladder carcinoma cell line (Krutovskikh et

    al., 1998). In a coculture experiment Esinduy et al. (1995)

    demonstrated inhibition of growth of neoplastically trans-

    formed, communication-deficient cells by cocultured, highlycommunicating, nontransformed cells. Not only do these stud-

    ies show potential for the use of connexins in gene therapy but

    the upregulation of GJIC by drugs or dietary factors may also

    provide chemoprevention (Trosko and Chang, 2001). Most

    important will be attention to the relative expression of differ-

    ent Cxs, the balance of Cx43:Cx32 being important in the GJIC

    of human malignant prostate epithelial cells (Carruba et al.,

    2002). A clear reduction of malignant potential of human

    hepatoma cells was indicated by an inhibition of dedifferenti-

    ation, reappearance of cell adhesion complex, and reduced

    proliferation coupled with enhanced apoptosis through trans-

    fection of Cx26 cDNA (Maramatsu et al., 2002; Yano and

    Yamasaki, 2001).Additional evidence for the role of Cxs as tumor suppressors

    has emerged from studies on transgenic mice lacking Cx genes.

    Temmeet al. (1997) reported increased susceptibility of Cx32

    knockout mice for the development of spontaneous and chem-

    ically induced liver tumors. Interestingly, phenobarbitone treat-

    ment did not further increase the incidence of liver tumors in

    Cx32-null animals, supporting the role of Cx32 as a crucial

    target in tumor promotion (Moennikeset al., 2000). Connexin

    expression is also inversely correlated to metastatic potential.

    The loss of cooperation between cells when GJ are disrupted

    may lead to heterogeneity and cell dissociation, thus support-

    ing metastasis (Carystinos et al., 2001). More specifically,

    Cx26 expression can reduce invasiveness of a human hepatoma

    cell line through E-cadherin expression and suppression of

    matrix metalloproteinase 9 (Yano and Yamasaki, 2001). It needs

    to be recognized, however, that different Cxs appear to have

    contrasting effects on metastagenicity (Carystinoset al.,2001).

    Role of GJIC in Bystander Effect in Tumor Therapy

    GJIC may also play an important role in cancer treatment in

    both radiotherapy and gene therapy. When radiotherapy is

    FIG. 2. A possible mechanism whereby inhibition of GJIC may contribute to tumor development.

    150 FORUM

  • 8/12/2019 Toxicol. Sci. 2003 Chipman 146 53

    6/8

    carried out, the possibility arises that positive effects may be

    seen in unirradiated areas of organs that may involve bystander

    effects mediated both by GJIC as well as through the release of

    signaling compounds into medium (Mothersill and Seymour,

    2001).

    Furthermore, transfection of a small number of tumor cells

    with the herpes simplex virus thymidine kinase gene (HSV-TK) followed by treatment with the nucleotide analogue gan-

    ciclovir (GCV) leads to phosphorylation of the latter and

    incorporation into the cells DNA, thus terminating DNA po-

    lymerization and leading to cell death. Tumors can enter re-

    gression even when only a small number of cells are treated,

    demonstrating bystander killing. This is thought to be GJIC-

    mediated, as it is believed that phosphorylated GCV may pass

    through gap junctions as with natural nucleotides. Such by-

    stander killing of tumor cells may be enhanced by transfection

    with connexin genes (Tanaka et al., 2001) provided that the

    function of their products can be maintained. Another use of

    connexins in gene therapy is focused on human glioblastoma.

    When cultured in vitro and transformed with cx43 (Huang etal., 2001) such cells become sensitive to chemotherapeutic

    drugs that induce apoptosis (i.e., etoposide, paditaxel, and

    doxorubicin).

    Conclusion

    The apparent role of the connexin family of proteins in such

    a wide range of cell functions relating to differentiation, pro-

    liferation, cell death, and migration, and the multitude of

    potential mechanisms of disruption, not surprisingly leads to

    the implication of these molecules in the processes of carcino-

    genesis and renders them potential useful targets in cancer

    therapy. A major future direction to aid understanding will bethrough targeted mutagenesis (Plum et al., 2000), dominant

    negative mutants (Yamasaki et al., 1999), and knock-in of

    specific cx genes in mice.

    REFERENCES

    Azzam, E. I., de Toledo, S. M., and Little, J. B. (2001). Direct evidence for the

    participation of gap junction-mediated intercellular communication in the

    transmission of damage signals from -particle irradiated to nonirradiated

    cells. Proc. Natl. Acad. Sci. U.S.A. 98, 473478.

    Bai, S., Schoenfeld, A., Pietrangelo, A., and Burk, R. D. (1995). Basal

    promoter of the rat connexin 32 gene: Identification and characterization of

    an essential element and its DNA-binding protein. Mol. Cell. Biol. 15,14391445.

    Brosnan, C. F., Scemes, E., and Spray, D. C. (2001). Cytokine regulation of

    gap-junction connectivity: An open-and-shut case or changing partners at

    the nexus? Am. J. Path. 158, 15651569.

    Bukauskas, F. F., Jordan, K., Bukauskiene, A., Bennet, M. V. L., Lampe, P. D.,

    Laird, D. W., and Verselis, V. K. (2000). Clustering of connexin 43-

    enhanced green fluorescent protein gap junction channels and functional

    coupling in living cells. Proc. Natl. Acad. Sci. U.S.A. 97, 25562561.

    Carruba, G., Stefano, R., Cocciadiferro, L., Saladino, F., Di Christina, A.,

    Tokar, E., Quader, S. T., Webber, M. M., and Castagnetta, L. (2002).

    Intercellular communication and human prostate carcinogenesis. Ann. N.Y.

    Acad. Sci. 963, 156168.

    Carystinos, G. D., Bier, A., and Batist, G. (2001). The role of connexin-

    mediated cellcell communication in breast cancer metastasis.J. Mammary

    Gland Biol. Neoplasia 6, 431440.

    Chen, S. C., Pelletier, D. B., Ao, P., and Boynton, A. L. (1995b). Connexin43

    reverses the phenotype of transformed cells and alters their expression of

    cyclin/cyclin-dependent kinases.Cell Growth Differ. 6, 681690.

    Chen, Z.-Q., Lefebvre, D., Bai, X.-H., Reaume, A., Rossant, J., and Lye, S. J.

    (1995a). Identification of two regulatory elements within the promoter

    region of the mouse connexin 43 gene. J. Biol. Chem. 270, 38633868.

    Cowles, C., Lee, A. J., and Chipman, J. K. (2000). Carbon tetrachloride

    reduces connexin 32 expression in rat liver in vivo only in association with

    hepatotoxicity.Toxicology 148, 64 (Abstract).

    Defamie, N., Mograbi, B., Roger, C., Cronier, L., Malassine, A., Brucker-

    Davis, F., Fenichel, P., Segretain, D., and Pointis, G. (2001). Disruption of

    gap junctional intercellular communication by lindane is associated with

    aberrant localization of connexin43 and zonula occludens-1 in 42GPA9

    Sertoli cells. Carcinogenesis 22, 15371542.

    Dermietzel, R., Yancey, S. B., Traub, O., Willecke, K., and Revel, J. P. (1987).

    Major loss of the 28-kD protein of gap junction in proliferating hepatocytes.

    J. Cell Biol.105,

    19251934.Duflot-Dancer, A., Mesnil, M., and Yamasaki, H. (1997). Dominant-negative

    abrogation of connexin-mediated cell growth control by mutant connexin

    genes.Oncogene 15, 21512158.

    Edwards, G. O., Meldrum, R., Wharton, C. W., Botchway, S. W., and Chip-

    man, J. K. (in press). Assessing the effect of ionizing radiation on a gap

    junction-mediated component of the bystander effect. Proceedings British

    Toxicology Society Meeting, April 2002. Toxicology.

    Eghbali, B., Kessler, J. A., Reid, L. M., Roy, C., and Spray, D. C. (1991).

    Involvement of gap junctions in tumorigenesis: Transfection of tumor cells

    with connexin 32 cDNA retards growthin vivo.Proc. Natl. Acad. Sci U.S.A.

    88, 1070110705.

    Elcock, F. J., Deag, E., Roberts, R. A., and Chipman, J. K. (2000). Nafenopin

    causes protein kinase C-mediated serine phosphorylation and loss of func-

    tion of connexin 32 protein in rat hepatocytes without aberrant expression or

    localisation. Toxicol. Sci. 56, 8694.

    Elmore, E., Milman, H., and Wyatt, G. (1987). Applications of the Chinese

    hamster V79 metabolic cooperation assay in toxicology.Adv. Mod. Environ.

    Toxicol. 14, 265292.

    Esinduy, C. B, Chang, C. C., Trosko, J. E., and Ruch, R. J. (1995). In vitro

    growth inhibition of neoplastically transformed cells by non-transformed

    cells: Requirement for gap junctional intercellular communication. Carci-

    nogenesis16, 915921.

    Fladmark, K. E., Gjertsen, B. T., Molven, A., Mellgren, G., Vintermyr, O. K.,

    and Doskeland, S. O. (1997). Gap junctions and growth control in liver

    regeneration and in isolated rat hepatocytes. Hepatology 25, 847855.

    Hernandez-Blazquez, F. J., Joazeiro, P. P., Omori, Y., and Yamasaki, H.

    (2001). Control of intracellular movement of connexins by E-cadherin in

    murine skin papilloma cells. Exp. Cell Res. 270, 235247.

    Holder, J. W., Elmore, E., and Barrett, J. C. (1993). Gap junction function andcancer. Cancer Res. 53, 34753485.

    Huang, R. P., Hossain, M. Z., Huang, R., Gano, J., Fan, Y., and Boynton

    (2001). Connexin 43 (Cx43) enhances chemotherapy-induced apoptosis in

    human glioblastoma cells. Int. J. Cancer92, 130138.

    Ito, S., Tsuda, M., Yoshitake, A., Yanai,T., and Masegi, T. (1998). Effect of

    phenobarbital on hepatic gap junctional intercellular communication in rats.

    Toxicol. Pathol. 26, 253259.

    Jou, Y. S., Layhe, B., Matesic, D. F., Chang, C. C., de Feijter, A. W.,

    Lockwood, L., Welsh, C. W., Klaunig, J. E., and Trosko, J. E. (1995).

    Inhibition of gap junctional intercellular communication and malignant

    151FORUM

  • 8/12/2019 Toxicol. Sci. 2003 Chipman 146 53

    7/8

    transformation of rat liver epithelial cells by neu oncogene. Carcinogenesis

    16,311317.

    Kasten, M. M., and Giordano, A. (1998). pRb and the cdks in apoptosis and the

    cell cycle. Cell Death Differ. 5, 132140.

    Kausalya, P. J., Reichert, M., and Hunziker, W. (2001). Connexin45 directly

    binds to ZO-1 and localizes to the tight junction region in epithelial MDCK

    cells. FEBS Lett. 505, 9296.

    King, L. K., and Cidlowski, J. A. (1995). Cell cycle and apoptosis: Common

    pathways of life and death. J. Cell. Biochem. 58, 175180.

    Klaunig, J. E., Ruch, R. J., and Lin, E. L. (1989). Effects of trichloroethylene

    and its metabolites on rodent hepatocyte intercellular communication. Toxi-

    col. Appl. Pharmacol. 99, 454465.

    Koffler, L., Roshong, S., Kyu Park, I., Cesen-Cummings, K., Thompson, D. C.,

    Dwyer-Nield, L. D., Rice, P., Mamay, C., Malkinson, A. M., and Ruch, R. J.

    (2000). Growth inhibition in G(1) and altered expression of cyclin D1 and

    p27(kip-1) after forced connexin expression in lung and liver carcinoma

    cells. J. Cell. Biochem. 79, 347354.

    Kojima, T., Sawada, N., Zhong, Y., Oyamada, M., and Mori, M. (1994).

    Sequential changes in intercellular junctions between hepatocytes during the

    course of acute liver injury and restoration after thioacetamide treatment.

    Virchows Arch. 425, 407412.

    Kolaja, K. L., Engelken, D. T., and Klaassen, C. D. (2000a). Inhibition of

    gap-junctional communication in intact rat liver by nongenotoxic hepato-

    carcinogens.Toxicology 146, 1522.

    Kolaja, K. L., Petrick, J. S., and Klaassen, C. D. (2000b). Inhibition of

    gap-junctional-intercellular communication in thyroid-follicular cells by

    propylthiouracil and low iodine diet. Toxicology 143, 195202.

    Kren, B. T., Kumar, N. M., Wang, S.-Q., Gilula, N. B., and Steer, C. J. (1993).

    Differential regulation of multiple gap junction transcripts and proteins

    during liver regeneration. J. Cell. Biol. 123, 707718.

    Krutovskikh, V. A., Mesnil, M., Mazzoleni, G., and Yamasaki, H. (1995).

    Inhibition of rat liver gap junction intercellular communication by tumor-

    promoting agentsin vivo. Association with aberrant localization of connexin

    proteins. Lab. Invest. 72, 571577.

    Krutovskikh, V. A., Yamasaki, H., Tusda, T., and Asamoto, M. (1998).

    Inhibition of intrinsic gap junction intercellular communication and en-

    hancement of tumorigenicity of the rat bladder carcinoma cell line BC31 by

    a dominant negative connexin 43 mutant. Mol. Carcinog. 23, 254261.

    Lampe, P. D., Kurata, W. E., Warn-Cramer, B. J., and Lau, A. F. (1998).

    Formation of a distinct connexin43 phosphoisoform in mitotic cells is

    dependent upon p34cdc2 kinase. J. Cell Sci. 111, 833841.

    Lefebvre, D. L., Piersanti, M., Bai, X.-H., Chen, Z.-Q., and Lye, S. J. (1995).

    Myometrial transcriptional regulation of the gap junction gene, connexin-43.

    Reprod. Fertil. Dev. 7, 603611.

    Lin, J. H., Weigel, H., Cotrina, M. L., Liu, S., Bueno, E., Hansen, A. J., Hansen

    Goldman, S., and Nedergaard, M. (1998). Gap junction-mediated propaga-

    tion and amplification of cell injury. Nat. Neurosci. 1, 494500.

    Loewenstein, W. R. (1981). Junctional intercellular communication: The cell-

    to-cell membrane channel. Physiol. Rev. 61, 829913.

    Loewenstein, W. R., and Rose, B. (1992). The cellcell channel in the control

    of growth. Semin. Cell Biol. 3, 5979.

    Mally, A., and Chipman, J. K. (2002). Non-genotoxic carcinogens: Early

    effects on gap junctions, cell proliferation, and apoptosis in the rat. Toxi-

    cology180, 233248.

    Maramatsu, A., Iwai, M., Morikawa, T., Tanaka, S., Mori, T., Harad, Y., and

    Okanoue, T. (2002). Influence of transfection with connexin 26 gene on

    malignant potential of human hepatoma cells. Carcinogenesis23, 351358.

    Mehta, P. P., Hotz-Wagenblatt, A., Rose, B., Shalloway, D., and Loewenstein,

    W. R. (1991). Incorporation of the gene for a cell-cell channel protein into

    transformed cells leads to normalization of growth. J. Membr. Biol. 124,

    207225.

    Mesnil, M., Fitzgerald, D. J., and Yamasaki, H. (1988). Phenobarbital specif-

    ically reduces gap junction protein mRNA level in rat liver. Mol. Carcinog.

    1, 7981.

    Moennikes, O., Buchmann, A., Romualdi, A., Ott, T., Werringloer, J., Wil-

    lecke, K., and Schwarz, M. (2000). Lack of phenobarbital-mediated promo-

    tion of hepatocarcinogenesis in connexin32-null mice. Cancer Res. 60,50875091.

    Moorby, C., and Patel, M. (2001). Dual functions for connexins: Cx43 regu-

    lates cell growth independently of gap junction formation. Exp. Cell Res.

    271, 238248.

    Mothersill, C., and Seymour, C. (2001). Radiation-induced bystander effects:

    Past history and future directions. Radiation Res. 155, 759767.

    Neuhaus, I. M., Dahl, G., and Werner, (1995). Use of alternate promoters for

    tissue-specific expression of the gene coding for connexin32. Gene 158,

    257262.

    Neveu, M. J., Hully, J. R., Paul, D. L., and Pitot, H. C. (1990). Reversible

    alteration in the expression of the gap junction protein Connexin 32 during

    tumor promotion in rat liver and its role during cell proliferation. Cancer

    Commun.2, 2131.

    Omori, Y., and Yamasaki, H. (1999). Gap-junction proteins connexin32 andconnexin43 partially acquire growth-suppressive function in HeLa cells by

    deletion of their C-terminal tails. Carcinogenesis 20, 19131918.

    Pandey, S., and Wang, E. (1995). Cells en route to apoptosis are characterized

    by the upregulation of c-fos, c-myc, c-jun, cdc2, and RB phosphorylation,

    resembling events of early cell-cycle traverse.J. Cell. Biochem58,135150.

    Peracchia, C., Wang, X. G., and Peracchia L. L. (2000). Chemical gating of

    gap junction channels. Methods 20, 188195.

    Piechocki, M. P., Burk, R. D., and Ruch, R. J. (1999). Regulation of con-

    nexin32 and connexin 43 gene expressions by DNA methylation in rat liver

    cells.Carcinogenesis 20, 401406.

    Piechocki, M. P, Toti, R. M., Fernstrom, M. J., Burk, R. D., and Ruch, R. J.

    (2000). Liver cell-specific transcriptional regulation of connexin32. Bio-

    chim. Biophys. Acta 149, 107122.

    Plante, I., Charbonneau, M., and Cyr, D. G. (2002). Decreased gap junctional

    intercellular communication in hexachlorobenzene-induced, gender-specific

    hepatic tumor formation in the rat. Carcinogenesis 23, 12431249.

    Plum, A., Hallas, G., Magin, T., Dombrowski, F., Hagendorff, A., Schuma-

    cher, B., Wolpert, C., Kim, J., Lamers, W. H., Evert, M., Meda, P., Traub,

    O., and Willecke, K. (2000). Unique and shared functions of different

    connexins in mice. Curr. Biol. 10, 10831091.

    Rae, R. S., Mehta, P. P., Chang, C. C., Trosko, J. E., and Ruch, R. J. (1998).

    Neoplastic phenotype of gap-junctional intercellular communication-defi-

    cient WB rat liver epithelial cells and its reversal by forced expression of

    connexin 32. Mol. Carcinog. 22, 120127.

    Rivedal, E., and Opsahl, H. (2001). Role of PKC and MAP kinase in EGF- and

    TPA-induced connexin43 phosphorylation and inhibition of gap junction

    intercellular communication in rat liver epithelial cells. Carcinogenesis 22,

    15431550.

    Rosenkranz, H. S., Pollack, N., and Cunningham, A. R. (1988). Nongenotoxicmechanisms in carcinogenesis: Role of inhibited intercellular communica-

    tion. InBanbury Report31.Carcinogen Risk Assessment: New Directions in

    the Quantitative and Qualitative Assessment Aspects (R. W. Hart and F. D.

    Hoerger, Eds.), pp 139170. Cold Spring Harbor Laboratory Press, Cold

    Spring Harbor, New York.

    Ruch, R. J., Guan, X., and Siegler, K. (1995). Inhibition of gap junctional

    intercellular communication and enhancement of growth in BALB/c 3T3

    cells treated with connexin43 antisense oligonucleotides.Mol. Carcinog.14,

    269274.

    Saez, J. C., Berthoud, V. M., Moreno, A. P. and Spray, D. C. (1993). Gap

    152 FORUM

  • 8/12/2019 Toxicol. Sci. 2003 Chipman 146 53

    8/8

    junctions: Multiplicity of controls in differentiated and undifferentiated cells

    and possible functional implications. Adv. Second Messenger Phosphopro-

    tein Res. 27, 163168.

    Saez, J. C., Connor, J. A., Spray, D. C., and Bennett, M. V. (1989). Hepatocyte

    gap junctions are permeable to the second messenger, inositol 1,4,5-

    trisphosphate, and to calcium ions. Proc. Natl. Acad. Sci. U.S.A. 86, 2708

    2712.

    Saez, J. C., Martinez, A. D., Branes, M. C., and Gonzalez, H. E. (1998).Regulation of gap junctions by protein phosphorylation. Braz. J. Med. Biol.

    Res. 31, 593600.

    Sugimura, T., and Ushijima, T. (2000). Genetic and epigenetic alterations in

    carcinogenesis.Mutat. Res. 462, 235246.

    Swierenga, S. H., and Yamasaki, H. (1992). Performance of tests for cell

    transformation and gap-junction intercellular communication for detecting

    nongenotoxic carcinogenic activity. IARC Sci. Publ. 1992 (116), 165193.

    Tan, L.-W., Bianco, T., and Dubrovic, A. (2002). Variable promoter region

    CpG island methylation of the putative tumor suppressor gene Connexin 26

    in breast cancer. Carcinogenesis 23, 231236.

    Tanaka, T., Yamasaki, H., and Mesnil, M. (2001). Stimulation of intercellular

    communication of poor-communicating cells by gap-junction-competent

    cells enhances the HSV-TK/GCV bystander effect in vitro. Int. J. Cancer

    91,538542.

    Tateno, C., Ito, S., Tanaka, M., Oyamada, M., and Yoshitake, A. (1994). Effect

    of DDT on hepatic gap junctional intercellular communication in rats.

    Carcinogenesis15, 517521.

    Temme, A., Buchmann, A., Gabriel, H. D., Nelles, E., Schwarz, M., and

    Willecke, K. (1997). High incidence of spontaneous and chemically induced

    liver tumors in mice deficient for connexin 32. Curr. Biol. 7, 713716.

    Temme, A., Ott, T., Dombrowski, F., and Willecke, K. (2000). The extent of

    synchronous initiation and termination of DNA synthesis in regenerating

    mouse liver is dependent on connexin32 expressing gap junctions.J. Hepa-

    tology 32, 627635.

    Trosko, J. E., and Chang, C.C. (2001) Mechanism of up-regulated gap junc-

    tional intercellular communication during chemoprevention and chemother-

    apy of cancer. Mutat. Res. 480481,219229.

    Trosko, J. E., Chang, C. C., and Upham, B. (2002). Modulation of gap

    junctional communication by epigenetic toxicants: A shared mechanism

    in teratogenesis, carcinogenesis, atherogenesis, immunomodulation, repro-

    ductive- and neurotoxicities. In Biomarkers of Environmental Associated

    Disease(S. H. Wilson and W. A. Suk, Eds.), pp. 445454. Lewis Publish-

    ers, Boca Raton, FL.

    Trosko, J. E., Chang, C. C., Upham, B., and Wilson, M. R. (1998). Epigenetic

    toxicology as toxicant-induced changes in intracellular signaling leading to

    altered gap junctional intercellular communication.Toxicol. Lett. 102103,

    7178.

    Trosko, J. E., and Goodman, J. I. (1994). Intercellular communication may

    facilitate apoptosis: Implications for tumor promotion. Mol. Carcinog. 11,

    812.

    Trosko, J. E., and Ruch, R. J. (1998). Role of cellcell communication in

    carcinogenesis.Front. Biosci. 3, D208236.

    Watanabe, H., and Tonosaki, A. (1995). Gap junctions in apoptosis. In

    Progress in Cell Research(Y. Kanno, K. Kataoka, Y. Shiba, Y. Shibata, and

    T. Shimazu, Eds.), Vol 4. Elsevier, Amsterdam.

    Willecke, E., Eiberger, J., Degen, J., Eckardt, D., Romualdi, A., Guldenagel,

    M., Deutsch, U., and Sohl, G. (2002). Structural and functional diversity of

    connexin genes in the mouse and human genome. Biol. Chem. 383, 752

    737.

    Wilson, R. M., Close, T. W., and Trosko, J. E. (2000). Cell population

    dynamics (apoptosis, mitosis, and cellcell communication) during disrup-

    tion of homeostasis. Exp. Cell Res. 254, 257268

    Yamasaki H. (1990). Gap junctional intercellular communication and carcino-

    genesis.Carcinogenesis, 11, 10511058.

    Yamasaki, H., Krutovskikh, V., Mesnil, M., Tanaka, T., Zaidan-Dagli, M. L.,

    and Omori, Y. (1999). Role of connexin (gap junction) genes in cell growth

    control and carcinogenesis. C. R. Acad. Sci. III322, 151159.

    Yano, T., Hernandez-Blazquez, F. J., Omori, Y., and Yamasaki, H. (2001).

    Reduction of malignant phenotype of HEPG2 cell is associated with the

    expression of connexin 26 but not connexin 32. Carcinogenesis 22, 1593

    1600.

    Yano, T., and Yamasaki, H. (2001). Regulation of cellular invasion and matrix

    metalloproteinase activity in HepG2 cell by connexin 26 transfection. Mol.

    Carcinog. 3, 101109.

    Yu, W., Dahl, G., and Werner, R. (1994) The connexin43 gene is responsive

    to oestrogen. Proc. R. Soc. Lond. [Biol.] 255, 125132.

    Zhang, Y. W., Morita, I., Ikeda, M., Ma, K. W., and Murota, S. (2001).

    Connexin43 suppresses proliferation of osteosarcoma U2OS cells throughpost-transcriptional regulation of p27. Oncogene 20, 41384149.

    Zhou, H., Randers-Pehrson, G., Waldren, C. A., Vannais, D., Hall, E. J., and

    Hei, T. K. (2000). Induction of a bystander mutagenic effect of -particles

    in mammalian cells. Proc. Natl. Acad. Sci. U.S.A. 97, 20992104.

    153FORUM