12-Lipoxygenases and 12(S)-HETE: role in cancer metastasis

32
Cancerand Metastasis'Reviews 13: 365-396, 1994. © 1994 KluwerAcademic Publishers. Printedin the Netherlands. 12-Lipoxygenases and 12(S)-HETE: role in cancer metastasis Kenneth V. Honn L'2'3'4,Dean G. Tang I, Xiang Gao ~, Igor A. Butovich 1'5,Bin Liu 1'¢', Jozsef Timar 1'7 and Wolfgang Hagmann ~'s Departments oF Radiation Oncology, 2 Chemistry, and 3 Pathology, Wayne State University, Detroit, MI48202; 4 Gershenson Radiation Oncology Center, Harper Hospital, Detroit, MI 48201; 5 Laboratory' for Chemical Enzymology, Institute of Bioorganic Chemistry, 1 Murmanskaya St., Kiev, 253660, Ukraine," 7 First Institute of Pathology and Experimental Cancer Research, Semmelweis Medical University, Budapest, Hungary; '~Deutsches Krebsforschungszentrum, Abt. Tumorbiochemie 0245, Im Neuenheimer Feld 280, D-69120 Heidelberg, Germany: ~ Present address: Department of Medicine, Division of Hematology/Oncology, Duke University, Durham, NC 27710 Key words: eicosanoids, 12-1ipoxygenase, metastasis, adhesion, 12(S)-HETE, protein kinase C Abstract Arachidonic acid metabolites have been implicated in multiple steps of carcinogenesis. Their role in tumor cell metastasis, the ultimate challenge for the treatment of cancer patients, are however not well-documented. Arachidonic acid is primarily metabolized through three pathways, i.e., cyclooxygenase, lipoxygenase, and P450-dependent monooxygenase. In this review we focus our attention on one specific lipoxygenase, i.e., 12-1ipoxygenase, and its potential role in modulating the metastatic process. In mammalian cells there exist three types of 12-1ipoxygenases which differ in tissue distribution, preferential substrates, and profile of their metabolites. Most of these 12-1ipoxygenases have been cloned and sequenced, and the molecular and bio- chemical determinants responsible for catalysis of specific substrates characterized. Solid tumor cells express 12-1ipoxygenase mRNA, possess 12-1ipoxygenase protein, and biosynthesize 12(S)-HETE [12(S)-hydroxyei- cosatetraenoic acid], as revealed by numerous experimental approaches. The ability of tumor cells to generate 12(S)-HETE is positively correlated to their metastatic potential. A large collection of experimental data suggest that 12(S)-HETE is a crucial intracellular signaling molecule that activates protein kinase C and medi- ates the biological functions of many growth factors and cytokines such as bFGF, PDGF, EGE and AME 12(S)-HETE plays a pivotal role in multiple steps of the metastatic 'cascade' encompassing tumor cell-vascu- lature interactions, tumor cell motility, proteolysis, invasion, and angiogenesis. The fact that 12-1ipoxygenase is expressed in a wide diversity of tumor cell lines and 12(S)-HETE is a key modulatory molecule in metastasis provides the rationale for targeting these molecules in anti-cancer and anti-metastasis therapeutic protocols. Introduction Metastasis is the hallmark of malignancy and the major obstacle to the cure of cancer patients. The overwhelming complexity of this pathological pro- cess is evidenced by our as-yet overall paucity of the understanding of the disease despite decades of re- search by cancer biologists and clinical oncologists. However, there is no doubt that significant progress has been made in deciphering many of the impor- tant molecular events associated with tumor pro- gression and dissemination. Cancer starts with cel- lular transformation in which germline gene (e.g., oncogenes and tumor suppressor genes) mutations with or without epigenetic deregulation render a normal cell tumorigenic. Transformed cells are

Transcript of 12-Lipoxygenases and 12(S)-HETE: role in cancer metastasis

Cancer and Metastasis' Reviews 13: 365-396, 1994. © 1994 Kluwer Academic Publishers. Printed in the Netherlands.

12-Lipoxygenases and 12(S)-HETE: role in cancer metastasis

Kenneth V. Honn L'2'3'4, Dean G. Tang I, Xiang Gao ~, Igor A. Butovich 1'5, Bin Liu 1'¢', Jozsef Timar 1'7 and Wolfgang Hagmann ~'s Departments o F Radiation Oncology, 2 Chemistry, and 3 Pathology, Wayne State University, Detroit, MI48202; 4 Gershenson Radiation Oncology Center, Harper Hospital, Detroit, MI 48201; 5 Laboratory' for Chemical Enzymology, Institute of Bioorganic Chemistry, 1 Murmanskaya St., Kiev, 253660, Ukraine," 7 First Institute of Pathology and Experimental Cancer Research, Semmelweis Medical University, Budapest, Hungary; '~ Deutsches Krebsforschungszentrum, Abt. Tumorbiochemie 0245, Im Neuenheimer Feld 280, D-69120 Heidelberg, Germany: ~ Present address: Department o f Medicine, Division of Hematology/Oncology, Duke University, Durham, NC 27710

Key words: eicosanoids, 12-1ipoxygenase, metastasis, adhesion, 12(S)-HETE, protein kinase C

Abstract

Arachidonic acid metabolites have been implicated in multiple steps of carcinogenesis. Their role in tumor cell metastasis, the ultimate challenge for the treatment of cancer patients, are however not well-documented. Arachidonic acid is primarily metabolized through three pathways, i.e., cyclooxygenase, lipoxygenase, and P450-dependent monooxygenase. In this review we focus our attention on one specific lipoxygenase, i.e., 12-1ipoxygenase, and its potential role in modulating the metastatic process. In mammalian cells there exist three types of 12-1ipoxygenases which differ in tissue distribution, preferential substrates, and profile of their metabolites. Most of these 12-1ipoxygenases have been cloned and sequenced, and the molecular and bio- chemical determinants responsible for catalysis of specific substrates characterized. Solid tumor cells express 12-1ipoxygenase mRNA, possess 12-1ipoxygenase protein, and biosynthesize 12(S)-HETE [12(S)-hydroxyei- cosatetraenoic acid], as revealed by numerous experimental approaches. The ability of tumor cells to generate 12(S)-HETE is positively correlated to their metastatic potential. A large collection of experimental data suggest that 12(S)-HETE is a crucial intracellular signaling molecule that activates protein kinase C and medi- ates the biological functions of many growth factors and cytokines such as bFGF, PDGF, EGE and AME 12(S)-HETE plays a pivotal role in multiple steps of the metastatic 'cascade' encompassing tumor cell-vascu- lature interactions, tumor cell motility, proteolysis, invasion, and angiogenesis. The fact that 12-1ipoxygenase is expressed in a wide diversity of tumor cell lines and 12(S)-HETE is a key modulatory molecule in metastasis provides the rationale for targeting these molecules in anti-cancer and anti-metastasis therapeutic protocols.

Introduction

Metastasis is the hallmark of malignancy and the major obstacle to the cure of cancer patients. The overwhelming complexity of this pathological pro- cess is evidenced by our as-yet overall paucity of the understanding of the disease despite decades of re- search by cancer biologists and clinical oncologists.

However, there is no doubt that significant progress has been made in deciphering many of the impor- tant molecular events associated with tumor pro- gression and dissemination. Cancer starts with cel- lular transformation in which germline gene (e.g., oncogenes and tumor suppressor genes) mutations with or without epigenetic deregulation render a normal cell tumorigenic. Transformed cells are

366

characterized by uncontrolled cell proliferation in- compatible with normal organ functions. Cell trans- formation itself, however, will not result in the gen- eration of metastatic tumor cell variants. Further genetic alterations and environmental insults are needed to trigger the dominant tumor cell clones that are able to seed themselves outside the primary lesion. For metastasizing tumor cells, completion of the 'metastatic cascade' is a tremendously challeng- ing endeavor, in which they must survive multiple 'checkpoints' such as vigilant host immune surveil- lance and harsh hemodynamic shear forces and overcome unfavorable biological barriers such as basement membrane. A small population of spreading tumor cells will eventually succeed colo- nizing a distant organ and forming metastases. Im- portant biological parameters in the metastatic pro- cess include angiogenesis, intravasation/extravasa- tion, proteolysis, locomotion/invasion, and cell pro- liferation [14]. These elements are not isolated but interdependent entities and are regulated by exter- nal driving forces exemplified by various growth factors, cytokines, chemokines, motility factors and a plethora of bioactive lipid molecules, among which are arachidonic acid (AA) metabolites.

AA is liberated from membrane phospholipids either by the action of phospholipase A 2 or through the concerted actions of phospholipase C and dia- cylglycerol lipase or of phospholipase D and phos- pholipase A 2 [6]. Once released, AA is metabolized through three major pathways. The cyclooxygenase pathway generates prostacyclin (PGI2), thrombox- ane (TXA2), and various prostaglandins; the lipoxy- genase pathway gives rise to a variety of hydroxyei- cosatetraenoic acids (HETEs), leukotrienes, and li- poxins; and the cytochrome P-450 monooxygenase pathway form various epoxyeicosatrienoic acids (EET) [for reviews see ref. 6-9]. Arachidonic acid and many of its metabolites have been implicated in multiple steps of tumorigenesis and metastasis [9- 13; also refer to relevant chapters in this issue]. The important role of eicosanoids in modulating cancer metastasis is best illustrated by prostacyclin/throm- boxane which in opposite directions regulate tumor cell-vessel wall interactions as well as metastasis [14-17; also see Chapter 11]. More recently, several lipoxygenase products (such as 12- and 15-HETE)

also have been demonstrated to be closely involved in regulating the metastatic process [18-21]. Since numerous comprehensive review articles exist in the literature documenting the general relationship between AA metabolites and tumor cell metastasis, we choose to focus our attention on only one branch of the lipoxygenase pathway, i.e., 12-1ipoxygenase pathway which generates as its major product 12(S)-HETE [12(S)-hydroxyeicosatetraenoic acid]. Our elaboration will start with the 12-1ipoxygenase gene structure and then move to the enzymology of the protein. Subsequently, the role of 12(S)-HETE in mediating tumor cell signal transduction and modulating various metastatic phenotypes is dis- cussed. Finally, the potential of using 12-1ipoxyge- nase as a prognostic marker for certain types of hu- man cancer and 12-1ipoxygenase inhibitors as anti- metastasis drugs is discussed.

12-Lipoxygenase: gene structure and regulation

Lipoxygenases insert molecular oxygen into cis, cis- pentadiene-containing lipid molecules generating various hydroperoxides. So far three animal lipoxy- genases have been identified; they are 5, 12-, and 15-1ipoxygenases named after the positions where oxygen is inserted to the AA moiety. 5-1ipoxyge- nase metabolizes AA to hydroperoxyeicosatetrae- noic acid (5-HPETE) which is further converted to 5-HETE and leukotrienes. 12-1ipoxygenase (EC 1.13.11.31) metabolizes AA to 12-HPETE which is subsequently transformed to 12-HETE and hepoxi- lins. Similarly, 15-1ipoxygenase catalyzes the reac- tion that generates 15-HPETE and subsequently 15- HETE and lipoxins.

Unlike 5- and 15-1ipoxygenases, 12-1ipoxygenase exists as isoforms within the same species [22]. Bio- chemical and immunological evidence indicates that 12-1ipoxygenases expressed in bovine leuko- cytes, platelets, and tracheal epithelium are dis- tinctly different enzymes [23, 24]. Information from cloned 12-1ipoxygenase cDNAs confirmed the ex- istence of different isoforms, cDNAs of 5-, 12- and 15-1ipoxygenase have been cloned from several dif- ferent species [25-33]. Introduction of human, por- cine, bovine or rat 12-1ipoxygenase cDNAs into eu-

karyotic cells or bacteria enables the transfectants to specifically metabolize AA to 12(S)-HETE [26, 31-33], although a small amount [less than 1/10 of the amount of 12(S)-HETE) of 15(S)-HETE also was generated in the case of bovine airway epithe- lial and rat brain 12-1ipoxygenases [26, 32]. Similar- ly, when cloned human reticulocyte 15-1ipoxyge- nase cDNA was transfected into bacteria, the cell lysate was able to convert AA predominantly to 15(S)-HETE and to a much lesser extent, to 12(S)- HETE [less than 1/10 of the amount of 15(S)- HETE)] [34].

Three types of 12-lipoxygenases have been re- ported. The first is human platelet-type 12-1ipoxy- genase expressed normally in platelets, HEL (hu- man erythroleukemia) cells, and umbilical vein en- dothelial cells [35]. Platelet-type 12-1ipoxygenase metabolizes only AA (but not C-18 fatty acids such as linoleic acid) to form exclusively 12(S)-HETE [36]. The second is porcine leukocyte-type 12-1ipox- ygenase which metabolizes both arachidonic acid and linoleic acid thus generating 12(S)-HETE as well as small amounts of 15(S)-HETE [37]. The third type of 12-1ipoxygenase (sometimes termed epithelial 12-1ipoxygenase) has been isolated from bovine tracheal epithelial cells [26] and rat brain [32], which shares more homology with 15-1ipoxy- genase and leukocyte-type 12-1ipoxygenase than with platelet-type 12-1ipoxygenase. This type of 12- lipoxygenase, like reticulocyte 15-1ipoxygenase and leukocyte-type 12-1ipoxygenase, catalyzes the for- mation of both 12(S)-HETE and 15(S)-HETE [26, 32, 38, 39]. Recently, mouse homologues of both platelet-type and leukocyte-type 12-1ipoxygenases have been reported [40].

When the three types of 12-1ipoxygenases are compared, 323 of the 661 amino acids are found to be identical (Fig. 1). Of these 323 amino acids, 47 residues differ from their corresponding sequences in 15-1ipoxygenase. In this set of 47 amino acids dif- ferences between 12- and 15-1ipoxygenases, 4 resid- ues are conserved between the two known 15-1ipox- ygenase sequences [27, 41]. When these 4 amino acids in human 15-1ipoxygenase are mutated to ami- no acids found in 12-1ipoxygenase, the enzyme cata- lyzes 12- and 15-oxygenation equipotently [42]. Fur- ther analysis of enzymes with single, double, triple

367

and quadruple mutations showed that Met418 is the primary determinant of positional specificity. When Met418 of 15-1ipoxygenase is replaced with a Val (the corresponding amino acid in 12-1ipoxygenase), the resulting enzyme (15-1ipoxygenase, M418V) performs 12- and 15-oxygenation equally thus gen- erating equal amounts of 12- and 15(S)-HETE [42]. No other mutation is found to be responsible for the positional specificity change. The region immedi- ately surrounding amino acid 418 is highly con- served in all mammalian lipoxygenases. Further mutagenesis studies demonstrated that the atoms of the side chain at position 418 are in contact with the substrate [42]. The replacement of the Met with a Val results in a 17% reduction of sidechain vol- ume. This would lead to a shift in substrate binding if the atoms of this side chain were to interact with the substrate. As Met and Val are both hydrophobic amino acids, they could participate in binding a fat- ty acid substrate. The site responsible for hydrogen abstraction of the wild-type enzyme is close to C-13 on AA, leading predominantly to 15-oxygenation. Presumably, the 15-1ipoxygenase M418V has a deeper substrate-binding pocket than the wild-type enzyme, so the C13 and C-10 of bound AA are equi- distant from the reaction center, thus the alignment between the methylenes and the hydrogen acceptor is imperfect, leading to equal 12- and 15-oxygena- tion. Thus M418 and its surrounding amino acids are responsible for the positional specificity of li- poxygenases.

When the cDNA and deduced amino acid se- quences of 5-, 12-, 15- and soybean lipoxygenases from different sources were compared, it was found that all of lipoxygenases contain a highly conserved sequence in the form of His-(X)4-His-(X)4-His-(X ) 17-His-(X)s-His which represents a potential bind- ing site for non-heme iron. This sequence is found at positions 363-400 for human 5-1ipoxygenase [29, 43], 355-392 for human 15-1ipoxygenase [27, 411 and human platelet 12-1ipoxygenase [31], 357-394 for porcine leukocyte 12-1ipoxygenase [25], and 356- 393 for bovine tracheal epithelium and rat brain 12- lipoxygenases [26, 32; Fig. 1]. Site-directed muta- genesis studies of these conserved histidine resid- ues in human 5-1ipoxygenase by two independent groups have shown that two of five histidine resid-

368

Hum pl Boy ep Pot ic Rat br Mou pl Mou ic

Hum pl Boy ep Pot ic Rat br Mou 1 Mou ~c

Hum pl Boy ep Por Ic Rat br

MOU

Hum pl Boy ep Pot ic Rat br Mou pl Mou ic

Hum pl Bov ep Por ic Rat br

Mou

Hum pl Boy ep Pot ic Rat br

Mou

Hum pl Bov ep Pot Ic Rat br

Mou

Hum p l Boy ep Pot IC Rat br Mou pl Mou ic

Hum pl Boy ep Pot Ic Rat br Mou 1 Mou ~c

Hum pl Bov ep Pot ic Rat br Mou pl Mou Ic

MGRYRIRVATGAWLF SGS YNRVQLWLVGT RGEAELE LQLRPARGEEEE FD HDVAEDLG- LLQFVRLP.KH 68 MGLYRVRVS TG S S F CAGS NNQVH LWLVGE HGEAALGWAVRPARGKEVEFQVDVS EYLGRLL - FVKLRKR 68 MGLYRVRVS TG S S F YAGS QNQVQLWLVGQHGEAALGWC LRPARGKE TEF S VDV SE YLGP LL- FVKLRKR 68 MGVYRIRVS TGD S K YAGSNNEVYLWLVGQ HGEAS LGK L LRPCRD SEAEFKVDV SE YLGP LL- FVRVQKW 68 MGAYRVRVVTGAWLF SGS LNLVRLWLVGE HREAKLE LQLRPARGKEEEFDFDVPED LGP L - QFVKLHKQ 68 MGVYRIRVS2~D S VYA~N~E]/YLWL I GQHGEAS LGK LFRP C~N SEA~KV~SEYLGPLL -~__VRVQKW 68

HWLVDDAWFCD RI TVQGP GACA- EVAFPCYRWVQGED I LSLPE GTARLPGDNALDMFQKHREKELKDRQ 136 HLLSDDAWFCNWI SVQGPGASGNEFRFPCYRWVEGDGI LSLPEGTGRTVVDDP QGLFKKHREEELAERR 137 HLLQDDAWFCNWI SVQGPGANGDEFRFPCYRWVEGDRI LSLPEGTARTVVDDP QGLFKKHREEELAERR 137 HYLTDDAWFCNWI SVKGP GDQGSEYMFPCYRWVQGRS I LSLPEGTGCTVVED SQGLFRKHREEELEERR 137 H TVVDDAWFCNL I TVQGP G- T SAE AVFP CYRWVQGEG I L SLP EGQARLAGDNALDVFQKYREKELKERQ 136 HYLKEDAWFCNWI S~KGPGDQGSEYTFPCYRWVQ~T S ILNLPEGTGCTVVED S QGLFRNH~EELEERR 137

Q I YCWATWKEGLP LT IAAD RKDDLPPNMRFHEEKRLDF EWTLKAGALEMALKRVYTLL S SWNC LEDFDQ 205 KLYRWGNWKDGL I LNIAGAT INDLPVDERFLEDKRIDFEASLTKGLAD LAIKD S LN ILTCWKSLDDFNR 206 KLYRWGNWKDGL I LNIAS TGI HD LPVDERF LEDKRIDFEASLAKGLAD LAVKD S LNVLMSWNSLD SFNR 206 SLYRWGNWKDGS I LNVAAAS I SDLPVDQRFREDKRI EFEASQV IGVMD TVVNFP INTVTCWKSLDDFNC 206 QTYCWATWKEGLP QT IAAD CKDD LPPNMRFHEEKRLDFEWT LKAGVLEMGLKRVYT LL RSWNH LEDFD Q 205 SLXR~GN~GT I LNV~AT S I SDLPVDQ~D~LEFEASQVI~TMD TVINFP KNTVTCWKSLDDFNY 206

IFW- GQKSALAEKVRQCWQDDELF SYQF LNGANPMLLRRSTSLP SRLVLP SGMEE LQAQLEKE LQNGS L 273 IFWCGQ - S KLAE RVRD SWKEDALFGYQF LNGTNPMLLRRSVRLP ARLE F P P GMGE LQAE LEKE LQQGT L 274 IFWCGQ- S KLAEQVRD SWKEDALFGYQF LNGTNPM_LLRH SVE LP ARLK F P P GMEELQAQLEKE LQGGT L 274 VF K SG - HT KMAERVRN SWKEDAFFGYQF LNGANPMVLKRSTC LPARLVF P P GME KLQAQLNKE LQKGT L 274 IFW- C-QKSALAEKVHQCWQEDELFGYQF LNGANPMLLRRST S LP SRLVLP SGMEELQAQLEKE LKNGS L 273 VF.K S~- H T KM~R~TRN SWKE~IAF E G y ~ L K R ~ T C LPA~VFP P ~E K~Q~EE LKKGT L 274

FEADF I LLDGI PANVIRGEKQYLAAP LVMLKME P NGKLQPMVI Q IQPP S P S sP TP TLFLP SDP P LAW-LL 342 FEADF SLLMGI KANVILCTQQYVAAPLVMLKLQPDGKLLPMAIQLQLP HKGSP PPPLFLPTDPPMTWLL 343 FEADFSLLDGIKANVI LC SQQYLAVPLVMLKLQPDGKLLPMVIQLQLPREGSP LPPLF LPTDPPMVWLL 343 FEADFFLLDGIKANVI LC SQQYLAAPLVMLKLMPDGQLLP IAIQLELP KTGSTPPP LFTPSDP PMDWLL 343 FEADFILLDGIPANVIRGEPQYLAAPLVMLRMDPGGKLLPMAIQIQPPNP S SPAPTLFLP SDPPLAWLL 342 FEADFFLLDGI KANV7 LC SQOYL~LQPDGQLLP I A/.QLE LPKTGST PPP IETELDPPMD~ 343

* * * * *

AKSWVRNSDFQLHE I QYH LLNTHLVAEVIAVATMRCLP GLHP I FKF P I P H IRYTME INTRARTQL I SDG 411 AKCWVRS SDFQLHELH SHLLRGHLVAEVXAVATMRCLP S IHPMFKLLIPHLRYTME IN IRARTGLVSD S 412 AKCWVRS SDFQLHELH SHLLRGHLMAEVIAVATMRCLP S IHP IFKLLIPHFRYTME INVRARNGLVSDL 412 AKCWVRS SDLQLHELQAHLLRGHLMAE LFAVATMRCLP SVHPVFKLLVPHLLYTME 7NVRARSDLI SER 412 AKIWVRNSDFQLQELQFHLLNTHLVAEVIAVATMRCLPGLHP IFKLLVPHIRYTME INTRTRTQLI SDG 411 AKCWVRS SD LQ~E LQA~RGH LVAEVFAVATMRCLP S VHP VFKL LVPHLLYTME INV~S DL ISE R 412

G I FDKAVS TGGGGHVQLLRRAAAQLTYC S - LCP PDDLADRGLLGLPGALYAHDALRLWE I IARYVEG IV 479 GVFDQWS TGGGGHVE LLRRAAALLTY S S - FCP PDDLADRGLLGVE S $ FYAQDALRLWEVI SRYVEG IV 480 GIFDQVVSTGGGGHVELLQRAGAFLTY- SGLLGVKS SFYSFCP PDD LADYAQDALRLWE I LSRYVEGIV 480 GFFDKAMS TGGGGH LD LLKQAGAF LTYC S - LCP PDDLAERGLLD I E TC FYAKDALRLWQ IMNRYVVGMF 480 GIFDQVVS TGGGGHVQLLT RAVAQLTYH S - LCP PDD LANRGLLR I P SALYARDALQLWEVTARYVKGMV 479 GFFD KVMS TGGGGH LDLLKQAGAF LTYS S - LCP PDDLAERGLLD I D T C FYAK~QLWQVMN~V~M F 480

HLFYQRDD IVKGDP ELQAWCRE I TEVGLCQAQDRGFPVSFQ SQ SQLCHFLTMCVFTCTAQHAAINQGQL 54 SLHYKTDE SVRDD IELQAWCRD I TE IGLLGAQDRGFPVTLQSKDQLCHFVTMC IFTCTGQHS S THLGQL 54~ SLHYKTDE SVKEDFELQAWCREFTE IGLLGAQDRGFPVS LQSKEQLCHFVTMC IFTCTGQHS SNHLGQL 548 NLHYKTDKAVQDDYELQSWCREI TD IGLQGAQDRGFPT SLQSRAQACYF I TMC IFTCTAQHS SVH LGQL 548 HLF YQ SDD IVRGD P E LQAWCRE I TEVGLCHAQDRGFPVSFQ SRAQLCHFLTMCVF TCTAQHAA INQGQL 548 DLYYKTDQA_~QDDYELOSWCQE ITE IGLQGAODRGFPT SLOSRA~ACHF I TMC IFTCTA~S S IH LGOL 548

DWYAWVPNAP CTMRMP PP TTKEDVTMATVMGSLPDVRQACLQMAI SWHL S RRQPDMVP LGHHKEKYF S G 617 DWYSWVPNAPCTMRLPPPTTK-DVTLEKVMATLPNFHQASLQMS ITWQLGRRQP IMVALGQHEEEYFSG 617 DWYTWVPNAPCTMRLPPP TTK-DATLETVMATLPNFHQASLQMS ITWQLGRCQPTMVALGQHEEEYFSG 617 DWFYWVPNAPCTMRLPPP TTKE- ATMEKLMATLPNPNQS TLQ INVVWLLGRRQAVMVP LGQHSEEHFP N 617 DWYGWVPNAPCTMRMPP P T SKDDVTME TVMGS LPDVQKACLQMT I TWH LGRLQPDMVP LGHHTEKYF SD 617 ~.~F YWVPN~CT~LPP P KTK-D A2ME K L~AT LPNPNQ S T LO I NVV~LLG~P, QAV~ LGQH SEE HF P N 617

P KP KAVLNQFRTDLE KLE KE I TARNEQLDWPYEYLKP S C I ENSV~ 663 PEPKAVLKKFREELAALEKD IE I RNAQLDWPYEYLRP S LVENSV~II 663 PGPKAVLTKFREELAALDKD I EVRNAKLALPYEYLRP S RVENSV~I I 663 PEAKAVLKKFREE LAALDKE I E I RNKSLD IPYEYLRP SMVENSV~II 663 PRTKAVLSQFQADLDNLEKEITARNEQLDLPYEYLKP S RIENS I ~II 663 PEAKAVLKKFREE LAALD~EIE I RNKSLD IPYEYLR~LVENSVA~ 663

Fig. 1. Deduced amino acid sequence alignment of 12-1ipoxygenases. Sequences which are identical to that of human platelet-type 12- lipoxygenase are highlighted. Conserved sequences among 6 currently characterized 12-1ipoxygenases are underlined. Conserved histi- dine residues are marked with an asterisk (*). Conserved isoleucine residues are boxed. Spaces are introduced to allow optimal align- ment. Hum: human, pl: platelet-type, Boy: bovine, ep: epithelial, Por: porcine, lc: leukocyte-type, br: brain, Mou: mouse.

ues are impor t an t for the catalytic activity of l ipoxy-

genases [44, 45]. A l though six 12-1ipoxygenase c D N A s have been

c loned f rom five species, its genomic s t ruc ture has only b e e n charac te r ized f rom human , pig, and mouse [35, 40, 46, 47]. All 12-1ipoxygenase genes isolated have 14 exons and the size of the exons, but not introns, are highly conserved a m o n g dif ferent

species. T h e human, porcine, and mouse p r o m o t e r sequences of 12-1ipoxygenase genes have been identif ied by different groups [35, 40, 46, 47], yet lit- tle is k n o w n abou t their regulat ion. Mult iple G C boxes were found in the 5' region, raising the possi- bility of 12-1ipoxygenase as a housekeep ing gene. Howeve r , this is not very likely since 12-1ipoxyge-

nase activity is inducible [33, 48-51].

12-Lipoxygenase Promoter Structures

369

Mouse leukocyte @ I GT repeat I GA repeat I [ ~ ~ ~nt

Mouse platelet -----IAAGG repeat I [ ~ [ ] ~ lnt

leukocyte AP-2 AP-2 - ln~t Porcine

Human platelet lnt

Fig. 2. Comparison of various 12-1ipoxygenase promoters from different species. The potential regulatory elements and repeat sequences are boxed. Mou: mouse; Por: porcine; Hum: human: Lc: leukocyte; Pl: ptatelet; Int (?): (potential) initiator site.

All 12-lipoxygenase promoters, except the mouse platelet-type 12-1ipoxygenase promoter, contains no typical TATA box. The human platelet-type 12- lipoxygenase promoter contains 4 putative GC box- es, 2 CACCC boxes, 3 AP-2 sites, and 1 glucocorti- cold-responsive element (GRE) [35, 46]. The por- cine 12-1ipoxygenase promoter possesses 9 GC box- es and 2 AP-2 sites [47]. In mouse, platelet-type 12- lipoxygenase promoter has 3 GC boxes, 1 TATA site, 1 TATA-like site, and 1 AP-2 site, whereas leu- kocyte-type 12-1ipoxygenase promoter contains 1 GC box, 1 TATA-like site, and 1 AP-1 site [40]. We have isolated the 12-1ipoxygenase promoter from a human colon carcinoma cell line (Clone A) and the sequence is nearly identical to the published human platelet-type 12-1ipoxygenase promoter sequence with a point mutation affecting one of the four GC boxes (X. Gao and K.V. Honn, unpublished data). Potential regulatory elements in various 12-1ipoxy- genase promoters are compared in Fig. 2. The GRE is activated by glucocorticoid, androgen, mineralo- corticoid and progesterone [52]. The AP-2 site is in- ducible by TPA, cAMP and retinoic acid [52]. The AP-1 site could be induced by TPA and other pro- tein kinase C (PKC) activators. Indeed, it has been shown that treatment of HEL cells with TPA in- creases 12-1ipoxygenase mRNA [50] and activity [33]. Furthermore, EGF, glucose, and angiotensin II have been demonstrated to increase 12-1ipoxyge- nase mRNA expression, stimulate 12-1ipoxygenase activity and therefore the production of HETE and HODE (i.e., hydroxyoctadecadienoic acids) [4'8, 49, 51]. Recently, we and others identified authentic

platelet-type 12-1ipoxygenase mRNA in Clone A (human colon adenocarcinoma) cells [53] and ker- atinocytes [55] and observed that 12-1ipoxygenase mRNA and protein were upregulated by 12-(S)- HETE treatment in clone A cells (Gao and Honn, unpublished data).

12-Lipoxygenase: chemical basis for its enzymatic activity

Lipoxygenases are iron-containing, non-berne en- zymes primarily catalyzing the aerobic oxidation of polyunsaturated fatty acids (PUFAs) containing a 1,4-cis, cis-pentadiene fragment yielding the corre- sponding hydroperoxides as the primary reaction products. It is of great interest to explore the kinet- ics and mechanism(s) of lipoxygenation reactions. The most widespread scheme of lipoxygenase clas- sification is based on the positional specificity of the enzyme toward AA. Depending on the position of the carbon atom being oxidized four major classes of lipoxygenases are discerned, nameley 5-, 8-, 12-, and 15-1ipoxygenases. The experiments carried out with a set of positional isomers of arachidonic acid [38] and linoleic acid [57] suggest that the substan- tial differences in the reaction rate and the set of products formed by different enzymes can be ex- plained by non-optimal alignment of the hydrogen acceptor of the enzyme reaction center and the dou- ble allylic methylene group of the substrate. In this connection, it will be significant to know which part of the PUFA molecule is responsible for its posi-

370

tioning in the enzyme's active center, i.e., the me- thyl moiety or the carboxylic group? The data pub- lished by Gardner et al. [58] indicate that it is the carboxylic moiety that plays a crucial role in the substrate orientation. Kuhn et al. [59] also proposed that the wheat lipoxygenase utilized the carboxylic group of the substrate for it to be oriented and aligned in a proper manner. There is, however, re- cent evidence that the m-end PUFAs might also play an important role in the substrate binding to the enzyme active center [42]. Therefore it seems likely that both parts of a PUFA molecule are es- sential for the substrate positioning and binding to the enzyme.

At least one Met residue that has been thought to be important in lipoxygenase catalysis is located in a near proximity to the substrate binding center [42, 6062]. It has been shown that the well-known self- inactivation of reticulocyte lipoxygenase is accom- panied by oxidation of a single Met to its sulfoxide [60]. Soybean lipoxygenase, when covalently mod- ified with iodoacetic acid or ~-bromstearic acid, but not with iodacetamide (additional evidence of the important role of carboxylic group in the substrate binding) lost its activity in a time- and dose-depend- ent manner, suggesting that the protonated carbox- yl group of iodacetic acid is essential for the inhib- itor positioning [61]. Taken together, these results indicate that the Met residue under consideration is located near the carboxyl-binding site of lipoxyge- nase and its catalytic center. The crucial role of the Met418 in positioning of PUFA in mammalian 12- and 15-1ipoxygenases was also postulated by Sloane et al. [42], as discussed in the preceding section.

All lipoxygenases so far reported are non-heme iron-containing enzymes. Four to five conserved histidines found in mammalian as well as plant li- poxygenases are thought to serve as ligands for the iron ion [63-66]. Usually Fe/apoenzyme ratio ap- proaches 1, depending on the source of the enzyme and methods used to extract and purify it. It is known that the iron in the enzyme's active center can be in high-spin ferric (Fe 3+) or ferrous (Fe 2+) forms. The redox state of the Fe 2+ can be easily changed by addition of micromolar range concen- trations of hydroperoxides of fatty acids [67], hy- drogen peroxide [68] or even, as proposed recently,

by oxygen [69]. Ferric iron undergoes reduction un- der the influence of some reductants, and specifical- ly by PUFAs themselves [70]. The most probable state of iron in plant cells is the ferrous form, imply- ing that the enzyme must be activated by an appro- priate oxidant to be able to oxidize PUFAs. In con- trast, porcine leukocyte 5- and 12-1ipoxygenases, at least partly, are extracted from the cells and further purified in the ferric form, as revealed by direct EPR experiments [71, 66]. Thus, it appears that they do not need a PUFA hydroperoxide to be convert- ed into the catalytically active form.

Considering the mechanism of aerobic lipoxyge- nase reaction and not going beyond the universal 1,4-cis, cis-pentadiene moiety one has to take into account the following scheme which has been pro- posed to describe the course of events in the en- zyme catalytic center:

LO-Fe(III) LO-Fe(II)

enzymatic, aerobic . .

R- (w) (m)

HOO anaerobic

aerobic ' Mixture of (R,S)-isomers R-R

(V)

The PUFA (I), which is stereospecifically deprot- onated by some enzyme-bound nucleophile B (-) to form an enzyme-bound carbon-centered free rad- ical (II), undergoes either i) stereoselective oxyg- enation giving the corresponding hydroperoxide (III), or, under anaerobic conditions, ii) subsequent dissociation (IV) and recombination resulting in the formation of dimers of PUFA (V). Sometimes, a side reaction leading to the formation of the race- mic mixture of (R,S)-hydroperoxides (VI) occurs as the result of dissociation of complex (II) in the pres- ence of molecular oxygen and further non-enzy- matic oxygenation of radical (IV).

Most lipoxygenases are capable of further trans-

formations of hydroperoxides. It has been shown [70, 72] that mammalian as well as plant lipoxyge- nases catalyze transformation of hydroperoxide 0II) to ketones (VII) (so-called 'anaerobic' reac- tion):

LO ) >o H20

III VII

and double oxygenation of tri- or tetraenoic acids [73]. It recently has been demonstrated that in the presence of certain proton donors lipoxygenases can possess hydroperoxidase activity [74], and more intriguingly, can oxidize a very unusual substrate, 12-keto-(9Z)-octadecaenoic acid methyl ester (VI- II) yielding the methyl ester of 9,12-diketo-(10E)- octadecaenoic acid (IX) [75]:

O

02

VIII IX

The enzyme activity toward 9(E), 12(Z)-octade- cadienoic acid, a rotational isomer of linoleic acid, is a well-known phenomenon reported by Funk et al.

[76]. Therefore, lipoxygenases may not require, for their enzymatic catalysis, the presence of double al- lylic methylenes in their substrates. It appears that most types of lipoxygenases can also catalyze the formation of epoxide (X) from the corresponding hydroperoxide (III) [63-65]:

OOH

II!

Lo ~

7 H20 x

In addition, 12- and 15-1ipoxygenases can also cleave their primary reaction products to give an al- kane (XI) and an aldehyde (XII) [77]:

371

LO H ~ > + Alkane

H20 O XII Xl

III

Finally, peroxidase-like activity of lipoxygenases has also been demonstrated, but appropriate pro- ton donors are required [74]. The mechanism of the enzymatic oxidation of esterified (e.g., incorporat- ed into phospholipids) PUFAs is intrinsically the same as has been described for free PUFAs.

12-Lipoxygenase: modulation of expression, activity, and intracellular distribution by extra- and intracellular factors

The intracellular 12-1ipoxygenase is subject to mod- ulation and regulation at various steps encompass- ing gene transcription, presence of active enzyme, and its inactivation. The susceptibility of 12-1ipoxy- genase to these modulatory actions varies between the enzymes from different cell types. Thus the con- tribution of any or all of such regulatory factors to the overall activity of 12-1ipoxygenase in a specific cell type deserves detailed examination.

Expression of the mRNA for 12-1ipoxygenase has been reported for human platelets [31, 33], leuko- cytes [33], endothelial cells [35], epidermal cells [78], and also for various human tumor cells includ- ing erythroleukemia (HEL) cells [33, 35], colon car- cinoma (Clone A) cells [53], epidermoid carcinoma A431 cells [48], rat Walker carcinosarcoma (W256), murine Lewis Lung carcinoma (3LL) and amela- notic melanoma (B16a) cells [53]. The presence of 12-1ipoxygenase mRNA, however, is not paralleled in all cases by the existence of a detectable amount of 12-1ipoxygenase protein or activity. In Clone A and cultured 3LL cells, 12-1ipoxygenase mRNA is detectable after amplification by RT-PCR [53], but the 12-1ipoxygenase activity or protein could not be demonstrated with certainty (Hagmann and Honn, unpublished observations). Thus, at least in some cell types one possibility for the regulation of 12-1i- poxygenase appears to exist at the post-transcrip- tional level. On the other hand, the expression of

372

12-1ipoxygenase mRNA in a specific cell type can be modulated. Up-regulation of 12-1ipoxygenase mRNA within 10-18 h is elicited in A431 cells by EGF [48]. In aortic smooth muscle cells, glucose and angiotensin II also cause a 2-3 fold increase in 12-1ipoxygenase mRNA expression [51]. In HEL cells, conflicting evidence was reported showing both up-regulation [50] and down-regulation [79] of 12-1ipoxygenase mRNA after treatment of the cells with phorbol ester TPA. The reason for this discre- pancy is yet unclear, but nevertheless stresses the notion that the 12-1ipoxygenase mRNA expression in these cells is subject to regulation by extracellular factors. Interestingly, an additional way to induce 12-1ipoxygenase mRNA expression apparently ex- ists in Clone A cells, where 12-HETE, generated by 12-1ipoxygenase activity and subsequent reduction, can enhance 12-1ipoxygenase mRNA expression within a few hours (Gao and Honn, unpublished observations), thus arguing for the existence of a possible positive feedback regulation of 12-1ipoxy- genase expression by the product of its own enzy- matic activity.

In cells containing enzymatically active 12-1ipox- ygenase, the intracellular distribution of this activ- ity varies between different cells and does not nec- essarily correspond with the respective distribution of the 12-1ipoxygenase protein. The subcellular dis- tribution of 12-1ipoxygenase activity ranges from a predominantly cytosolic compartmentation in pla- telets, leukocytes, and epithelial cells [24, 80, 81] to an exclusive or preferential membrane-associated activity in A431, HEL, freshly isolated 3LL, and ep- ithelial cells [49, 79, 82-85]. In accordance with their inducing effect on 12-1ipoxygenase mRNA expres- sion, extracellular factors such as glucose, angioten- sin II, and EGF also cause an increase in 12-1ipoxy- genase activity in the respective cells or subcellular fractions [49, 51]. As to microsomal membranes from HEL cells, conflicting data document either an increase or a decrease of 12-1ipoxygenase activity following TPA treatment of cells [33, 79, 84]. On the other hand, deprivation of EGF-responsive A431 cells of this growth factor rapidly results in the loss of 12-1ipoxygenase activity, whereas re-addition of EGF or serum to such starved cells immediately re- stores the 12-1ipoxygenase activity in some cells

[189]. Interestingly, 12-1ipoxygenase was reported to be differentially sensitive to activating or inhib- itory factors depending on its intracellular localiza- tion. Glutathione (GSH) concentrations above 1.5 mM inhibit cytosolic 12-1ipoxygenase activity in platelets, whereas the membrane-associated 12-1i- poxygenase activity in HEL or ovine epithelial cells is already inhibited by 1 mM or 0.1 mM GSH, re- spectively [83, 84]. Analogously, dithiothreitol (DTT) inhibits the membrane-associated 12-1ipox- ygenase activity from HEL cells [196]. Inversely, hy- droperoxides such as 13-HPODE do not influence membrane-associated 12-1ipoxygenase activity [79, 83], but have been demonstrated to be sufficient for activating cytosolic 12-1ipoxygenase activity which is inactive in the absence of 13-HPODE [84].

The subcellular distribution of the 12-1ipoxyge- nase protein in some tumor cells does not parallel the intracellular localization of their 12-1ipoxyge- nase activity. In 3LL cells, the 12-1ipoxygenase pro- tein is localized predominantly in the cytosol [53, 85], as is their 12-1ipoxygenase activity [82, 85]. In A431, HEL, W256, and B16a cells, however, the predominantly cytosolic 12-1ipoxygenase protein is enzymatically far less active than the minor amount of membrane-associated 12-1ipoxygenase protein [53, 85] suggesting a tight regulation of 12-1ipoxyge- nase activity in the cytosol of these cells. In addi- tion, the subcellular distribution of 12-1ipoxygenase activity and protein can change upon stimulation of cells, suggesting a translocation of the enzyme pro- tein as a physiologic event during its activation pro- cess in cells [84-86]. Translocation of an enzyme from cytosol to membranes during its activation is known to occur with many enzymes such as PKC [87], PLC, PLA 2 [88, 92], diacyl- and 2-monoacyl- glycerol lipases [89], and 5-1ipoxygenase [90], whereas the reported 8-1ipoxygenase in murine skin apparently does not translocate to membranes in a Ca2+-dependent manner [91]. In platelets, such a Ca2+-induced stimulation was shown to result in an increase in membrane-associated 12-1ipoxygenase activity and a concurrent loss of its activity from the cytosol [86]. This Ca 2+-, thrombin-inducible shift in the subcellular localization of 12-1ipoxygenase ac- tivity [86] is paralleled by a concommitant translo- cation of the 12-1ipoxygenase protein (Hagmann

and Honn, unpublished observations), and the membrane-associated 12-1ipoxygenase can be reco- vered as fully active enzyme by decreasing the free Ca 2÷ concentration [86]. In contrast to platelets, a Ca2+-induced loss of cytosolic 12-1ipoxygenase ac- tivity in 3LL cells is reflected by translocation of 12- lipoxygenase protein, but this process results in a membrane-associated enzyme which is not fully ac- tive or becomes inactivated shortly after transloca- tion to the membrane site [85]. In HEL cells, Ca 2+ also induces a concentration-dependent decrease in the residual cytosolic 12-1ipoxygenase activity and a concomitant translocation of activity to membranes [84].

The discrepancy in some cells between distribu- tion of 12-1ipoxygenase protein and its activity in subcellular compartments suggests that the enzyme is inhibited in the cytosolic compartment. Such an inhibitory effect by cytosolic constituents on 12-1i- poxygenase activity has been suggested previously [49, 83]. These earlier studies, however, reported a cytosol-dependent decrease in specific activity rather than in total or relative 12-1ipoxygenase ac- tivity disregarding the fact that the addition of cyto- sol to the membrane fraction affects markedly the specific activity by increasing the protein content of the sample tested. However, in HEL cells addition of cytosol to the enzymatically active membrane fraction does not alter its relative 12-1ipoxygenase activity, whereas its specific 12-1ipoxygenase activ- ity decreases dose-dependently with added cytosol [84].

12(S)-HETE: involvement in multiple steps of cancer metastasis

12(S)-HETE is the major AA metabolite of 12-1i- poxygenase. Many different types of normal cells including platelets, neutrophils, macrophages, en- dothelial cells, and smooth muscle cells can synthe- size 12(S)-HETE [7, 51, 158]. The physiological functions of this eicosanoid are not fully character- ized, but accumulated data indicate that it is in- volved in a wide-spectrum of biological activities such as stimulating insulin secretion by pancreatic tissue, suppressing renin production, chemoattract-

373

ing leukocytes, and facilitating the attachment of macrophages to rat glomeruli during inflammation (reviewed in Ref. 7). 12(S)-HETE also has been ob- served to reduce prostacyclin biosynthesis by vascu- lar endothelial cells [93] and play a vital role in pla- telet activation and aggregation [94]. More recently, 12(S)-HETE is found to be the most prominent AA metabolite in menstrual blood [95] and in intraute- rine tissues [96], however, the biological signifi- cance of these findings is not yet clear.

As discussed in the previous section, a variety of tumor cells including solid tumor cells express 12- lipoxygenase mRNA and protein [48, 49, 53, 84, 85, 87, 97]. RT-PCR together with immunoblotting re- vealed that cultured solid tumor cells from human, rat, and mouse express platelet-type 12-1ipoxyge- nase mRNA and protein [53, 84, 85, 87]. Reverse phase HPLC identified 12(S)-HETE as the major metabolite of AA in these tumor cells which is con- firmed by chiral phase HPLC and GC/MS analysis [97, 98]. Several lines of evidence suggest that the ability of tumor cells to synthesize 12(S)-HETE is a strong correlate of their metastatic potential. First, subpopulations of B16 amelanotic melanoma (B16a) have been isolated by centrifugal elutriation [99] that demonstrate differential metastatic capa- bilities. The low metastatic B16a (i.e., LM180) cells biosynthesized similar amounts of 12(S)-HETE and 5-HETE, while high metastatic B16a cells (i.e., HM340) generated predominantly 12(S)-HETE with only small quantities of 15-, 11-, and 5-HETEs [97]. Furthermore, HM340 cells synthesized 4 times more 12(S)-HETE than LM180 cells when equal amounts of substrates were supplied [97]. The gen- eration of higher amount of 12(S)-HETE in HM340 cells appears to result from the presence of a higher level of 12-1ipoxygenase mRNA in these cells [195]. Second, the correlation of 12(S)-HETE production and metastatic potential also was evaluated in sev- eral other tumor cell systems, i.e., Dunning rat pros- tate carcinoma (AT2.1 and GP 9F3, low metastatic cell lines; MAT Lu and MLL, high metastatic lines), murine B16 melanoma (F1, low metastatic; F10, high metastatic), and murine K-1735 melanoma (C1-11, low metastatic clone; M1, high metastatic clone). In all these experiments it was observed that the high metastatic tumor cell lines generate signif-

374

icantly higher level of 12(S)-HETE than low meta- static counterparts [195]. Third, tumor cell adhesion to fibronectin provokes a spike of 12(S)-HETE gen- eration within 10 rain indicating a late-type signal- ing activation of tumor cell 12-1ipoxygenase [112]. Morphological studies demonstrated that tumor cell spreading on fibronectin is followed by 12-1i- poxygenase translocation to the cell surface [112]. Fourth, adhesion of tumor cells to vascular endo- thelium is accompanied or immediately followed by a surge of 12(S)-HETE biosynthesis by tumor cells, which was correlated with tumor cell-induced en- dothelial cell retraction [98, 100]. LM180 B16a cells generated little 12(S)-HETE upon adhesion and did not induce endothelial cell retraction. In con- trast, HM340 B16a cells adhering to endothelium biosynthesized large amounts of 12(S)-HETE and induced prominent retraction of endothelial cell monolayers [98]. Fifth, pretreatment of tumor cells with a select platelet-type 12-1ipoxygenase inhib- itor, BHPP (N-benzyl-N-hydroxy-5-phenylpenta- namide; 53) ~ . . . . d0se-dependently inhibited adhesion- induced tumor cell biosynthesis of 12(S)-HETE and endothelial cell retraction [98], platelet-enhanced tumor cell-induced endothelial cell retraction [100], tumor cell (i.e., HM340 B16a cells) adhesion to en- dothelium and matrix [53, 97], and lung coloniza- tion by HM340 B16a cells [97]. Taken together, these data implicate 12-1ipoxygenase and 12(S)- HETE production as important determining par- ameters of tumor cell metastasis. In the subsequent discussions, we will analyze in detail the versatile modulatory role of 12(S)-HETE in various interme- diate steps of metastasis.

A. Effect of l2(S)-HETE on tumor cell interactions with extracellular matrix (ECM)

Tumor cell - ECM interactions are mediated by ad- hesion receptors such as integrins, non-integrins and proteoglycans where integrins are considered to be the predominant matrix receptors [18, 101, 102]. Tumor cell - ECM interactions are key and multiple events during tumor progression; i.e. in the release of tumor cells from the primary site, in the locomotion and invasion of dissociated tumor cells

in interstitium, in intravasation and extravasation, and in the eventual establishment of the metastatic nodule. It is conceivable that the regulation of the surface expression of these matrix receptors has a great impact on tumor cell adhesive potentials.

Integrin receptors, after synthesis, are stored in tumor cells as well as in normal cells intracellularly in so called adhesosomes [103] and tumor cells have a large pool of intracellular receptors [104]. How- ever, in normal cells, freshly synthesized and glyco- sylated integrins are not competent for their full range of ligand binding and signaling functions. Ad- hesion process initiated by either soluble or solid surface-bound ligands triggers cell activation. Li- gand binding to the surface exposed integrins in- duces conformational change of the receptors in- creasing the binding potential and modulating their binding specificity. In parallel, such physical alter- ations in the receptors induce signal transmission to the cell interior by activation of a receptor associ- ated tyrosine kinase, FAK (focal adhesion kinase; 105-107). As a consequence, cytoskeletal proteins become associated with the cytoplasmic domains of the integrins and this process immobilizes and con- centrates the receptors to the areas of cell - matrix contacts called focal adhesion plaques. During the cytoskeleton-dependent process the cell shape changes dramatically from round to spread.

Several extracellular factors, examplified by ECM-derived soluble ligands and ECM-bound cy- tokines (characteristically these are heparin-bind- ing cytokines such as TGF-[~, FGFs, VEGE etc.) may regulate or alter the adhesive properties of tu- mor cells. In the case of tumor cell - vessel wall in- teractions other cellular participants such as plate- lets and endothelial cells modulate the adhesion process by releasing cytokines and metabolites of AA to the micromilieu of the tumor cells. There- fore, a possibility is raised that the metabolites of AA generated in the extracellular space of tumor cells may modulate the tumor adhesiveness and thus tumor cell adhesion to ECM. Systematic stud- ies indicated that 12(S)-HETE is able to stimulate metastatic murine tumor cell adhesion to fibronec- tin and subendothelial matrix in a dose and time- dependent fashion, with a maximal effect observed at 0.1 ~M of 12(S)-HETE 15 min after stimulation

[104, 108]. Analysis of the mechanism of the 12(S)- HETE enhanced tumor cell adhesion indicated an increased surface expression of integrin receptors cdIb133 after 12-(S)-HETE treatment [104, 108]. In the murine tumor cell lines studied (3LL and B16a) the increased adhesiveness was due to overexpres- sion of cdIb133 cytoadhesin [109,110], a promiscuous matrix receptor physiologically expressed by plate- lets and cells of the hematopoietic lineage. Interest- ingly, TPA, a direct activator of PKC, mimicked the effect of 12(S)-HETE [108] suggesting a PKC-medi- ated pathway for the exogenous 12(S)-HETE effect (see below).

The 12(S)-HETE effect on tumor cell integrin ex- pression was not due to an increased transcription of the gene or increased protein translation. Rather it was due to a rapid receptor translocation to the cell surface from the intracellular pool [104, 111]. These studies further indicated that the 12(S)- HETE effect on integrin upregulation and en- hanced adhesion both depend on rnicrofilaments and intermediate filaments [104, 111].

12(S)-HETE does not only stimulate tumor cell adhesion to fibronectin but also promotes tumor cell spreading on this matrix protein, by inducing reestablishment of the filamentous cytoskeleton system and enhancing the formation of focal adhe- sion plaques [112]. This process appears to be de- pendent on the function of PKC [112]. The dose and time-frame of the 12(S)-HETE effect on spreading followed those observed previously for adhesion [112].

B. Effect o f exogenous 12(S)-HETE on tumor cell motility

Tumor cell motility is regulated by para- and auto- crine cytokines including SF (scatter factor), MF (motility factor) and AMF (autocrine motility fac- tor) [113]. These cytokines act through their corre- sponding signaling receptors expressed at the cell surface of tumor cells. AMF was isolated as the first autocrine motility cytokine produced by tumor cells [114]. A receptor for AMF, gp78, was later identified and sequenced [115]. These studies revealed a trans-

375

membrane protein with considerable homology to p53 [115].

Hydroxy fatty acid metabolites of AA including 12-(S)-HETE have been shown to be involved in chemotactic movement of leukocytes where they serve as chemoattractants [7]. Therefore there ex- ists a possibility that 12-(S)-HETE also may mod- ulate tumor cell movement. Studies on murine mel- anoma cell lines with different metastatic potential (B16a, K1735-C1.11 and K1735 M1) indicated that 12(S)-HETE induces tumor cell motility compara- ble to AMF itself [116]. Subsequent mechanistic studies demonstrated that 12-(S)-HETE, similar to its effect on tumor cell integrin receptor expression, increases gp78 expression at the cell surface in a dose and time-dependent manner [116]. The in- creased gp78 expression is due to translocation of the receptor to the cell surface from an intracellular tubulovesicular (TVS) structure positive for lysoso- real markers [117]. Studies also indicated that this TVS pool is closely associated with microtubules and intermediate filaments, therefore it is reason- able to assume that the increased AMF receptor ex- pression after 12(S)-HETE stimulation is also a cy- toskeleton-dependent process.

More recently, 12(S)-HETE also was observed to increase the invasive potential of AT2.1 rat prostate tumor cells [116, 118]. The tumor cell motility (as as- sessed by phagokinetic track assay) and invasion (as assessed by Matrigel invasion assay; Fig. 3) were significantly augmented by 12(S)-HETE [118]. The 12(S)-HETE effect was dramatically inhibited by a selective PKC inhibitor calphostin C as well as by a cell membrane-permeable Ca2+-chelator BAPTA [118], suggesting the involvement of PKC (see be- low). Whether the 12(S)-HETE-promoted motility involves gp78 and whether 12(S)-HETE-augment- ed invasion in the rat tumor cell system involves ac- tivation of proteolytic enzymes are at present un- clear.

The reciprocal relationship between 12-1ipoxyge- nase/12(S)-HETE and AMF/gp78 appears to be far more complicated than presented above. 12(S)- HETE apparently promotes an AMF-comparable and gp78-mediated tumor cell motility. On the other hand, gp78 also has been observed to stim- ulate 12(S)-HETE production in highly metastatic

376

Fig. 3. 12(S)-HETE enhances the invasive capacity of rat prostate carcinoma cells. The invasion assay was performed using the Biocoat Matrigel invasion chambers (Collaborative Res., Bedford, MA). Fifty thousand of rat AT2.1 prostate adenocarcinoma cells treated with either ethanol (a) or 0.1 pM 12(S)-H ETE (b) were placed into the upper chamber in 200 pl of serum free RPM11640 medium. To the lower chamber either RPMI or 3T3 cell conditioned medium was added. The cell invasion was quantitated as described [118]. It is obvious that 12(S)-HETE treatment (b) resulted in a significant increase in the invasion of AT2.1 cells, x 400.

murine melanoma cells [116]. More interestingly, AMF, which stimulated the motility of the high- (M1) but not low-metastatic (C1.11) variants of the K1735 murine melanoma, increased expression of 12-1ipoxygenase protein in M1 sublines exclusively. Also, only the M1 (but not C1.11) cells responded to AMF stimulation with increased endogenous 12(S)- HETE production and 12(S)-HETE, like AME only increased the motility of M1 cells [190]. Taken together, these data suggest that 12-1ipoxygenase and 12(S)-HETE may be important intermediate signaling elements in AMF-gp78 mediated tumor cell motility.

C. Effect of exogenous 12(S)-HETE on the release of lysosomal enzymes

Earlier studies on various normal cells indicated that 12(S)-HETE induces lysosomal enzyme or hor- mone secretion [7]. Furthermore, lipoxygenase me- tabolites have been shown to be involved in collage- nase IV release from tumor cells (reviewed in ref. 119). Secretion of proteins involves a so-called se- cretory mechanism, where the glycosylated protein is transported to the plasma membrane by the se- cretory pathway involving transport vesicles. Lyso- somal enzymes utilize a mechanism which directs these proteins to lysosomes by recognizing the M6P residues on the protein (M6P-R). Lysosomal en- zymes are released from these vesicles after intra- or extracellular signaling. Tumor cells produce a wide range of proteolytic enzymes by which they di- gest the surrounding ECM. These degradative en- zymes include collagenase IV, interstitial collage- nases, heparinases, urokinase and cathepsins with variable substrate specificities [120-122]. Secretion of these enzymes from tumor cells provides a means by which tumor cells can cross tissue boundaries. In normal cells, the majority of such proteolytic en- zymes are stored intracellularly and released only after appropriate stimulation. However, in tumor cells lysosomal enzymes, especially cathepsins, were found to be associated constitutively with the cell surface [121-123].

It is known that the translocation of lysosomal proteins is mediated by cytoskeletal elements, pri-

377

marily by microtubules. The demonstrated involve- ment of 12(S)-HETE in the regulation of receptor (such as integrin cdIb[33) translocations through cy- toskeleton-dependent mechanisms prompted us to study its effect on the surface expression and re- lease of cathepsin B from tumor cells. At a concen- tration which is optimal for increased adhesiveness and motility of tumor cells (i.e., 0.1 gM), 12(S)- HETE was observed to induce cathepsin B release from highly malignant cells, i.e., B16a murine mela- noma cells and MCF10AneoT hmnan mammary carcinoma cells [119, 122, 124]. Exogenous 12(S)- HETE first triggers the release of mature cathepsin B and then later the immature form, suggesting that the mature form of the enzyme is at or near the cell surface while the immature form is stored intracel- lularly. Subsequent morphologic studies indicated that, following 12(S)-HETE stimulation, the intra- cellular cathepsin B-containing vesicles are direct- ed to specialized cell surface areas, i.e., lamellipodia and filopodia [119]. Analysis of the release process induced by 12-(S)-HETE indicated that, similarly to its effect on tumor cell adhesion and motility, the stimulated release of cathepsin B is also a PKC-de- pendent process [119].

D. Effect of exogenous 12(S)-HETE on the tumor cell infrastructure and cytoskeleton

It is clear from the above discussions that 12(S)- HETE has a regulatory effect on complex cellular processes associated with cytoskeletal functions, such as adhesion to and spreading on ECM pro- teins, motility, lysosomal enzyme release and recep- tor translocation to the cell surface. Therefore, we have analyzed the effect of exogenous 12(S)-HETE on cytoskeleton in tumor cells [125, 126]. These studies indicated that 12(S)-HETE induces translo- cation of organelles from the perinuclear space to the cell periphery paralleled by a reversible rear- rangement of cytoskeletal filaments [125]. The ear- liest event post 12(S)-HETE treatment is a de- creased labeling of cytoplasmic stress fibers with a concomitantly enhanced cortical actin labeling. These alterations in microfilaments are accompa- nied or immediately followed by microtubule poly-

378

ermization and vimentin intermediate filament bundling and also by rearrangements of the focal adhesion protein vinculin [125, 126]. The 12(S)- HETE effect on tumor cell cytoskeleton is PKC-de- pendent since pretreatment of tumor cells with a se- lective PKC inhibitor, i.e., calphostin C abolished the 12(S)-HETE effects [125]. Interestingly, 12(S)- HETE-induced cytoskeletal alterations generally abate by ~ 30 min. However, simultaneous treat- ment of tumor cells with 12(S)-HETE and okadaic acid, a general serine/threonine protein phospha- tase inhibitor, prevented disappearance of the 12(S)-HETE effects by 30 min [126]. These obser- vations thus raise the possibility that a phosphatase (s), activated due to either the direct effect of 12(S)- HETE or homeostatic mechanisms, is responsible for dampening the 12(S)-HETE-triggered, PKC- dependent, and phosphorylation-mediated cytos- keleton rearrangement.

E. Effect of12 (S)-HETE on vascular endothelial cells: induction of endothelial cell retraction and modulation of integrin expression and cell proliferation

Tumor cell - endothelial cell interactions constitute one of the most important factors in the organ pref- erence of metastasis. Specific interactions of metas- tasizing tumor cells with end organ endothelial cells will, to a large degree, dictate the route of metasta- sis. In the vasculature, migrating tumor cells, uti- lizing cell surface adhesion molecules, first attach loosely to the intact endothelial cell monolayers by recognizing counterpart adhesion receptors on the latter cell type. This initial cell-cell adhesion leads to, by certain unknown mechanisms, the activation of tumor cells and/or endothelial cells (or other host cells such as platelets). Cellular activations result in an enhanced surface expression of adhesion recep- tors and/or a functional maturation in the binding competence of these adhesion molecules, leading to a tighter bonding of tumor cells to endothelium. This whole process of tumor cell - endothelium in- teractions has previously been analyzed at the molecular level with the 'docking and locking' hy- pothesis [18]. The hypothesis dissects tumor cell in-

teractions with target organ endothelial cells into two basic interdependent but distinct steps, as illus- trated in Fig. 4. The early interaction, i.e., the 'dock- ing', occurs between the 'rolling' tumor cells and underlying endothelial cells and generally is medi- ated by relatively weak and transient adhesion mechanisms involving carbohydrate-carbohydrate (such as Le×-Le x and glycosphingolipid-glycosphin- golipid; 191-192), carbohydrate-selectin (such as sLeX-E selectin), and protein-protein (such as PE- CAM-1-PECAM-I; 145) recognition (Fig. 4; see ref. 18 for detailed account). The initial tumor cell-en- dothelial cell interactions result in the activation of either cell type or other host cells such as platelets, which leads to increased surface expression and/or functional 'maturation' of other adhesion mole- cules, typically integrins, and finally to the second phase, tighter tumor cell adhesion, i.e., the 'locking' step (Fig. 4). The transition from the 'docking' stage to the 'locking' stage relies on the generation from both tumor cells and host cells (i.e., leukocytes, pla- telets, endothelial cells, etc) of various bioactive mediators including cytokines, chemokines, auto- crine factors, as well as lipid molecules examplified by 12(S)-HETE (Fig. 4; 18 and references therein).

After tumor cells have established tight, physical connections with endothelial monolayers, tumor cells spread and then exit from the vessel by induc- ing endothelial cell retraction. Although some types of tumor cells may extravasate by intravascu- lar proliferation followed by rupture of the vessel wall, the majority of experimental (morphological as well as functional) studies both in vivo [127] and in vitro [128, 129] have demonstrated that vascular endothelial cells retract prior to the extravasation of most tumor cell types. Therefore, tumor cell in- duced endothelial cell retraction has been proposed to facilitate tumor cell metastasis. Unfortunately, we are still far from understanding the molecular mechanisms underlying tumor cell-induced endo- thelial cell retraction. Although many factors in- cluding vasoactive substances (histamine, bradyki- nin, serotonin, etc), lipid mediators (LTC4, PAF, etc), thrombin, TNF-c~, oxygen radicals, neutral proteases, endotoxins, and 5'-radiation have been shown to increase endothelial cell permeability and induce physical endothelial cell retraction [130-

"Docking" Transition "Locking" 379

EC

*Weak and transient binding

*Mediated mainly by carbohydrate-carbo- hydrate recognition

*Activation of tumor cells and host cells (platelets, leukocytes, endothelial cells, etc)

*Generation of inflammatory cytokines, chemokines, and bioactive lipids

*Expression of inducible adhe- sion molecules such as selectins, ICAM, and VCAM

*Tighter tumor cell adhesion to endothelial cells

*Mediated principally by integrin receptors

*Integrins' are the major signal transducers for the subsequent molecular events (endothelial cell retraction, interaction with ECM, proteolysis, cell locomotion and invasion

Fig. 4. The 'docking and locking' hypothesis. Tumor cell interactions with vascular endothelial cells can be dissected into an initial phase of weak adhesion (i.e., 'docking') and a later phase of tight adhesion (i.e., 'locking'). Differential involvement of specific subsets of adhesion molecules in these two phase have been demonstrated (see text and ref. 18 for details). The transition from the 'docking' phase to the 'locking' stage is mediated via a large array of bioactive mediators including 12(S)-HETE. TC, tumor cell; EC, endothelial cells.

132], the biological mediators responsible for tumor cell induced endothelial cell retraction largely re- main elusive.

Five years ago, our laboratory presented the first experimental evidence that 12(S)-HETE, a biolog- ically relevant molecule derived from AA metabo- lism, is capable of inducing a non-destructive, re- versible retraction of cultured endothelial cell monolayers [133]. 12(S)-HETE induced endothelial cell retraction resulted in an enhanced tumor cell adhesion to exposed subendothelial matrix [133] and could be inhibited or blocked by prostacyclin or by 13(S)-HODE [134], reminescent of the bidirec- tional control of tumor cell alIb[33 integrin expres- sion by 12(S)-HETE/13(S)-HODE [108] and of tu- mor cell - platelet interactions by thromboxane A2 and prostacyclin [14-17]. Subsequent studies estab- lished that 12(S)-HETE also induces retraction of microvessel endothelial cells [135-137]. More im- portantly, we recently demonstrated that tumor cell adhesion to microvessel endothelium induced a non-denuding retraction of endothelial cell mono- layers [98], which, in many aspects, mimicked the 12(S)-HETE-induced endothelial cell retraction. Tumor cell induced endothelial cell retraction can be potentiated by homologous platelets [100], thus

providing an experimental rationale for the platelet contribution to tumor cell metastasis. Further in- vestigations revealed that both platelet-augmented and tumor cell induced endothelial cell retraction were mediated through 12(S)-HETE [98, 100]. These observations provide strong evidence for the possible scenario that in v ivo 12(S)-HETE is pro- duced at high concentrations in the localized areas of tumor cell-platelet-endothelial cell interactions. This de n o v o synthesized 12(S)-HETE, most prob- ably by tumor cells and more importantly from pla- telets, may be transferred to endothelial cells through transcellular metabolism to induce endo- thelial cell retraction, since there has been experi- mental evidence that small molecules can be trans- ferred from tumor cells to endothelial cells [138].

Endothelial cell retraction can be affected by a variety of mechanisms dependent on different caus- ative agents. Thus, histamine-induced endothelial cell retraction is mediated by ATP-dependent mi- crofilament rearrangement [139]. By contrast, thrombin-induced endothelial cell gap-formation is dependent on the levels of intracellular cAMP [140]. Recently, myosin light chain kinase has been implicated in endothelial cell retraction caused by many substances [141]. Our previous work demon-

380

strated that 12(S)-HETE is an activator of PKC in tumor cells [87]. Recently, we have observed that endothelial cell retraction induced by 12(S)-HETE is also a PKC-dependent process [135-137, 142]. 12(S)-HETE activates PKC in endothelial cells, which subsequently phosphorylates several major cytoskeletal elements including myosin light chain, actin and vimentin, leading to a reorganization of the cytoskeleton network (microfilament, interme- diate filament, and focal adhesions) and apparent endothelial cell retraction.

12(S)-HETE induced endothelial cell retraction may also involve alterations in adhesion molecules that normally are concentrated at cell-cell junc- tions. Specically, integrin av[33 and PECAM-1 (pla- telet endothelial cell adhesion molecule-l), an im- munoglobulin supergene family member that is ex- pressed on platelets, vascular endothelial cells, cer- tain subtypes of neutrophils, as well as in some solid tumor cells [143-145], are localized to the cell junc- tional areas in confluent endothelial cell cultures. 12(S)-HETE treatment results in a series of well- defined changes in c~v133-containing focal adhe- sions, leading to an eventual decrease in av[~3 local- ization to focal adhesion plaques at both cell body and cell-cell borders [135,136]. In contrast to its ef- fect on integrin o~vl33,12(S)-HETE does not appear to alter the normal cellular distribution patterns of another integrin receptor, i.e., c~5131, an authentic fi- bronectin receptor. The 12(S)-HETE induced reor- ganization of c~v133-enriched focal adhesions resem- bles its effect on vinculin and again appears to be dependent on PKC activation [135,136]. PECAM-1 has been proposed to be involved in maintaining the integrity of endothelial cell monolayers [143]. Treatment of post-confluent endothelial cell cultur- es with 12(S)-HETE dispersed cell junction-local- ized PECAM-1 molecules to the whole cell body (Tang and Honn, unpublished observations). Whether this alteration of PECAM-1 is a trigger or a consequence of 12(S)-HETE induced endothelial cell retraction remains to be established.

Interestingly, 12(S)-HETE also modulates the surface expression of integrin receptors on endo- thelial cells. 12(S)-HETE stimulation upregulates the surface expression of integrin av[33 (but not o~5131) in both large vessel [146] and microvessel

[136] endothelial cells. The increased czv[33 surface expression appears to be concomitant with a de- crease in ~v[33-enriched focal adhesions [136] and is a posttranscriptional, PKC- and cytoskeleton-de- pendent process [147]. The end result of 12(S)- HETE enhanced surface expression of (zvl]3 inte- grin receptors is increased tumor cell adhesion to the activated endothelium [136, 146, 147].

12(S)-HETE may as well be involved in regulat- ing endothelial cell growth. Early work demonstrat- ed that the DNA synthesis as well as proliferation of cultured fetal bovine aortic endothelial cells were inhibited by nordihydroguaiaretic acid (NDGA), a general lipoxygenase inhibitor, but not affected by the cyclooxygenase inhibitor indomethacin [149], suggesting lipoxygenase metabolites as endoge- nous regulators of endothelial cell growth. These endothelial cells generate as their major lipoxyge- nase metabolite 15(S)-HETE, with a minor amount of 12(S)-HETE [165]. Application of either of these two lipoxygenase metabolites promoted significant DNA synthesis as well as proliferation of endothe- lial cells [165]. Very recently, the vital role of lipoxy- genase products, 12(S)-HETE in particular, in con- trolling endothelial cell growth has been establish- ed in bovine capillary endothelial cells [166]. Beta FGF-, PDGF-, and serum-stimulated vascular cell (including both endothelial cells and smooth mus- cle cells) growth is dose-dependently inhibited by NDGA and a 12-1ipoxygenase inhibitor baicalein but not by indomethacin and a 5-1ipoxygenase in- hibitor 5,6-dehydroarachidonic acid [166]. The growth of unstimulated endothelial cells was also inhibited by NDGA and baicalein [166], further lending support to the concept that 12(S)-HETE and perhaps some other lipoxygenase metabolites such as 15(S)-HETE are necessary growth signals for vascular cells. PDGF and bFGF are well-known angiogenic factors. The observations that their cell growth-promoting effects are to certain extent me- diated through lipoxygenase metabolites such as 12(S)-HETE [166] strongly suggest that the latter by themselves may be important physiological an- giogenic factors. In fact, 15(S)-HETE has been shown to enhance capillary endothelial cell migra- tion at 0.1 gM and to be proangiogenic in an in vivo

rabbit corneal pocket assay (reviewed in ref. 165).

Whether 12(S)-HETE which is a structural isomer of 15(S)-HETE also shares similar proangiogenic properties is currently unknown. Considering that human stimulated blood can produce ~ 21 btM and 1 btM of 12(S)- and 15(S)-HETE, respectively [165] and that the levels of these metabolites can be fur- ther increased at the focal areas of tumor cell-plate- let-endothelial cell interactions, it is reasonable to assume that these metabolites, especially 12(S)- HETE, may play a crucial role in modulating vari- ous phenotypic properties of endothelial cells such as adhesiveness, migration, and proliferation, as well as in their responses to growth factors and cyto- kines.

E The mechanisms o faction of l2(S)-HETE: a physiological PKC activator

12-HETE is the most abundant AA metabolite from human platelets [148] and is produced by pla- telets of all mammalian species so far examined [6]. As presented above, some tumor cells also express 12-1ipoxygenase mRNA, possess 12-1ipoxygenase enzyme, and generate 12-HETE. Arachidonic acid metabolites, including various lipoxygenase prod- ucts, have been implicated in various signal trans- duction pathways involving growth factors, G-pro- teins, 1321 raS and ras-GAR cytokines, DAG and PKC, and neurotransmitters [149-158, 172, 173]. A growing body of recent evidence suggests that 12(S)-HETE may act as a second messenger in stim- ulus-response coupling in some cells [159-162]. For example, angiotensin II stimulates 12-HETE bio- synthesis and aldosterone secretion by adrenal cor- tical cells; pharmacologic inhibitors of 12-HETE biosynthesis suppress angiotensin II-induced aldos- terone secretion; and exogenous 12-HETE increas- es aldosterone secretion [160]. A similar series of observations suggest the participation of 12-HETE or its precursor in neurotransmitter peptide-in- duced hyperpolarization by Aplysia neuronal cells [159]. 12(S)-HETE, like LTB 4, has been shown to promote epidermal proliferation [161]. On the other hand, treatment of rat renal cortical slices with 12-HPETE or 12-HETE, but not LTB 4 or 5- HPETE, blocked the PGI 2- or iloprost-induced re-

381

nin secretion [163]. In addition, 12-HETE has been shown to significantly increase intracellular levels of cyclic AMP (cAMP) in human arterial smooth muscle cells [164].

Our laboratory has performed comprehensive studies on the molecular mechanisms underlying the versatile 12(S)-HETE effects on various tumor cells as well as vascular endothelial cells. Early ex- perimental observations that 12(S)-HETE mim- icked the tumor promoter PMA in enhancing tu- mor cell integrin expression and adhesion [108] lent credence to the postulate that 12(S)-HETE may work via activating PKC. Supportive evidence for this hypothesis was obtained when we observed that, in rat Walker carcinosarcoma (W256) cells, 12(S)-HETE induced a 100% increase in mem- brane-associated PKC activity and that down-regu- lation of PKC by prolonged treatment with PMA abolished the 12(S)-HETE enhanced W256 cell ad- hesion to endothelium [87]. Subsequently, it was observed that essentially all of the 12(S)-HETE evoked biological responses (in both tumor cells and endothelial cells) encompassing rearrange- ment of cytoskeleton, promotion of cell adhesion and spreading, induction of endothelial cell retrac- tion, release of proteolytic enzymes, and enhance- ment of cell motility can be abrogated by selective PKC inhibitors [15-19, 87, 112, 116, 118, 119, 125,126, 135-137, 147]. Experiments by others also have shown that 12(S)-HETE and some other HETEs (such as 15-HETE) activate PKC [174]. PKC has been considered one of the most important phos- pholipid/Ca2+-dependent serine/threonine protein kinases localized in the center of various signaling pathways [167]. It is now known that PKC is a family of protein kinase which has at least 10 members, some of which are Ca2+-dependent (so-called con- ventional PKCs or cPKC) and some Ca2+-independ - ent (including novel PKCs or nPKC and atypical PKCs or aPKC). Different tissues and cells differ- entially express a specific repertoire of PKC iso- forms that perform specialized functions. Although the precise role of PKC in multi-step tumorigenesis and cancer metastasis remains to be established, generally it is acknowledged that abnormal expres- sion of specific PKC isoforms is closely associated

382

198 -

120 -

8 8 -

7 0 -

5 6 -

3 8 -

3 2 -

1 2 3 4 5 6 7 8 9 10 11 12 13 14

- - - - ~aD .---" e m

Fig. 5. Murine B16a melanoma cells express only the a isoform of PKC. Whole cell lysates were prepared from either cultured B16a cells (lanes 1, 3, 5, 7, 9, 11, and 13) or rat brains (lanes 2, 4, 6, 8, 10, 12, and 14). Equal amounts of proteins were loaded onto 10% denaturing SDS-PAGE and separated proteins transferred to nitrocellulose. The membrane was then probed with isoform-specific anti-PKC pep- tide antibodies (kindly provided by Dr. Y. Hannun, Duke University). The M,W. (in Kd) standards were shown on the left.

with cellular t ransformat ion and progress ion [168-

171]. The wel l -known physiological act ivators of P K C

are diacylglycerol ( D A G ) and Ca 2+. H o w 12(S)-

H E T E activates P K C has not been decisively de-

fined. It appears that 12 (S ) -HETE in different cell

types activates different isoform(s) of PKC. In mu- rine B16a m e l a n o m a cells, PKCo~is the only i soform

that can be detec ted by immunoblo t t ing using iso-

form-specif ic ant i-pept ide antibodies (Fig. 5).

12 (S) -HETE st imulat ion leads to an increased

t ranslocat ion of P K C f rom the diffuse cytoplasmic

pool to the plasma m e m b r a n e (Fig. 6A and B).

W h e n B16a cells plated on f ibronect in were stim-

ulated with 12(S) -HETE, enhanced membrane-as - sociated P K C was observed to colocalize with vin-

A PKC B Vinculin

Fig. 7. B16a cells plated on fibronectin were stimulated with 0.1 gM of 12(S)-HETE for 15 min and then processed for double labeling of PKC (A) and vinculin (B). Bar, 10 gm.

culin to the focal adhesion plaques (Fig. 7). In rat AT2.1 prostate carcinoma cells, two isoforms of PKC, i.e., c~ and 5, were identified [118]. Interesting- ly, 12(S)-HETE preferentially activates PKCct by increasing the membrane association of this iso- form, resulting in an upregulated tumor cell motil- ity and invasion [118]. Treating tumor cells with thy- melea toxin, which is a selective activator of PKCcz over Ca2+-independent PKCS, also promoted tu- mor cell motility and invasion [118; Fig. 8], thus con- firming the preferential activation of PKCc~ by 12(S)-HETE in this tumor cell line. Further, BAP- TA, a Ca2+-chelator, dose-dependently inhibited 12(S)-HETE-promoted tumor cell motility and in- vasion [118; Fig. 9], suggesting that the PKC isoform activated by 12(S)-HETE is Ca2+-dependent. In mu- rine microvessel endothelial cells (i.e., CD3 cells), at least two isoforms of PKC are expressed (Liu, Tang, and Honn, unpublished observations). Which isoform of PKC or whether both isoforms of PKC are activated by 12(S)-HETE is not yet known. In rat adrenal glomerulosa cells, the c~ and ~ isoforms of PKC are expressed [175]. In sharp contrast, as de- scribed above, to the situation in tumor cells, 12(S)- HETE and 15(S)-HETE alike predominantly acti- vate PKCe, although angiotensin II increases the membrane-bound levels of both PKCc~ and PKCe [175].

Exactly how 12(S)-HETE activates PKC is un- clear. What is known is that this fatty acid does not directly activate PKC as assessed by PKC activity assays with 12(S)-HETE included in the reaction mixtures [193]. However, 12(S)-HETE stimulation of tumor cells is observed to be followed by a rapid accumulation of DAG and IP 3 [193], therefore sug- gesting PIP 2 metabolism through the action of PLC. To support this, blocking the functions of G pro- teins by pertussis toxin in B16a melanoma cells as well as inhibiting the PLC activity by specific antag- onists also suppress the effects of exogenous 12(S)- H E T E [193], indicating that 12(S)-HETE may func- tion via activating an upstream G-protein. Another apparent alternative for the accumulation of DAG following 12(S)-HETE treatment is through down- regulating the DAG kinase activity, as proposed by Setty et al. [165]. In the phosphoinositide cycle, DAG derived from various sources undergoes rap-

383

PKC

e- O

tla b- ttl n -

|

Fig. 6. 12(S)-HETE promotes PKCo~ association with plasma membrane, which is blocked by 13(S)-HODE. Immunofluores- cent labeling was performed using anti-PKCc~ as the primary an- tibody followed by incubating cells with FITC-labeled secondary antibody. In control (unstimulated) B16a cells, PKC takes the form of diffuse cytoplasmic staining (A). Stimulation of cells with 0.1 p.M of 12(S)-HETE (15' at 37 ° C) resulted in an in- creased association of PKC with the membrane (revealed as brightening dots in the picture, B). However, pretreatment of cells with 0.1/aM 13(S)-HODE (15" at 37 ° C) completely blocked the 12(S)-HETE effect (C). Bar, 10 btm.

384

"S

4.-a ¢--

O ( o

g,e

> , .t.-a

. ~

O

7-

"S

C- O (D

O

>

03 ~3 > d -

200

150

t00

50

0

250 B

200

150

100

50

-r- i

0

0 0.01 0.1

Thymelea Toxin (pM) #@

#@

T I I

T

I I

0.001 0,01 0.1

Thymelea toxin (txM)

A

Fig. 8. PKCc~ activator thymelea toxin promotes motility and invasiveness of rat prostate carcinoma cells (AT2.1). In motility assays (A),

cells were treated with 0.01 or 0.1 btM thymelea toxin for 60 min. Cells in the control group were treated with vehicle ethanol (0.002%).

Data are f rom a representat ive exper iment and similar results were obtained in two other experiments. *: p < 0.001 compared to control; +: p < 0.005 compared to 0.01 gM thymelea toxin. B. To determine the effect of thymelea toxin on AT2.I cell invasion, cells were mixed with toxin containing medium to the indicated final concentrat ions and loaded into the invasion chambers. Data represent mean + SEM

of three exper iments performed in duplicate. #: p < 0.05 compared to control; @: p < 0.05 compared to 0.001 btM thymelea toxin. For

detailed experimental protocols, see ref. 118.

250

200

150

O L

g

7~

5-

A

o k -

g £D

u? U~

el)

2> f -

100

50

0

i

350

300

250

200

150

100

50

0

+ 0.1 laM 12(S) -HETE A

I l

385

/ 1 5 25 0

BAPTA/AM (ktM) 5 25

+ 0.1 jaM 1 2 (5 ) -HETE

I I

TIT 1 5 12 .5

T

0 1 5 12.5

BAPTA/AM (~M)

B

Fig. 9. Effect of calcium chelator BAP T A on AT2.1 cell motility and invasion. In motility assays (A)~ cells were treated with indicated

concentrat ions of B A P T A / A M for 30 min or pretreated for 30 min followed by I2(S)-HETE (0.1 [aM) for 15 rain. Control cells were

treated with appropriate amount of vehicle DM S O (0.19 %). No difference in basal motility was observed when compared to cells treated with RPMI alone. Results are expressed as a percentage of the motility of control cells and are the mean + SEM of three separate

experiments. B. To determine the effect of B AP T A on invasion, AT2.1 cells (1 x 10 ~) were pretreated with indicated concentrat ion of

B A P T A / A M for l0 min, pelleted and resuspended in medium. Treated cells (5 x 104) were loaded into the invasion chambers and 12(S)~

H E T E or vehicle ethanol (0.03%) was added to the final concentrat ion of 0.1 ~aM. Data represent mean + SEM of four experiments.

id phosphorylation by DAG kinase to phosphatidic acid [176]. In bovine aorotic endothelial cells, 12(S)- HETE as well as 15(S)-HETE were found to inhibit the DAG kinase activity, leading to an accumula- tion of DAG mass in these cells [165]. Certainly it is possible that both mechanisms (i.e., the activation of upstream G-protein and PLC and the inhibition

of downstream DAG kinase) exist in cells resulting, eventually, in an increase in DAG and then activa- tion of specific PKC isoform(s).

In various experimental settings we observed that the 12(S)-HETE effects (integrin expression, adhesion, spreading, retraction, etc) on tumor cells and vascular endothelial cells are antagonized by

386

13(S)-HODE 12(S)-HETE

P / T u r n o v e r - - - - - - -

Inactive PKC 1

Protein phosphorylation and activation of other kinases

Adhesion, spreading, motility

+ Metastasis

Fig. 10. A schematic illustration of the molecular mechanisms of the 12(S)-HETE effects. Detailed description and relevant references are given in the text.

13(S)-HODE, a bioactive lipid derived from lino- leic acid metabolism by 15-1ipoxygenase [87, 99,108, 134, 146]. The 13(S)-HODE effects appear to be mediated through an inhibition of PKC activation by 12(S)-HETE. However, as in the case of 12(S)- HETE, 13(S)-HODE does not directly suppress the PKC activation [193]. In epidermis where 13- HODE is the major metabolite of linoleic acid, this fatty acid was found to be incorporated into 13- HODE-containing diacylglycerol, i.e., 13-HODE- DAG [177]. Interestingly, 13-HODE-DAG was demonstrated to selectively inhibit the PKC-[3 ac- tivity in epidermal cells [177]. Nevertheless, wheth- er 13-HODE-DAG directly binds to PKC-[3 and subsequently inhibits its activity or 13-HODE-

DAG achieves the inhibition of PKC via some other signaling intermediates is unknown.

It is now well-appreciated that various cyclooxy- genase products including prostaglandins, prosta- cyclin, and thromboxanes exert their functions through binding to specific cell surface receptors [22, 178]. So far many receptors and subtype recep- tors have been cloned and sequenced for these eico- sanoids and most of them are shown to be typical G protein-linked receptors [22, 178-181]. Accumulat- ing experimental data suggest that 12(S)-HETE al- so may have a cognate cell surface receptor or bind- ing site(s). Binding studies using human epidermal cell line SCL-II demonstrated that these cells pos- sess a single class of binding sites for 12(S)-HETE with a Kd of 2.6 nM and a Bma X of 216,000 sites per

cell [182]. In human platelets, 12-HETE has previ- ously been shown to block PGH2/TXA2-induced platelet aggregation [183]. Later, part of the inhib- itory effect of 12(S)-HETE on PGH2/TXA2-in- duced platelet aggregation was found to be due to an inhibition of PGHR/TXA 2 binding to their com- mon receptor sites [184]. These observations sug- gest a possibility of 12(S)-HETE interaction with or binding to the PGH2/TXA z receptors. More recent- ly, 12(S)-HETE binding sites have been reported in Lewis lung carcinoma (LLC) cells [185]. Intriguing- ly, subsequent subcellular fractionation studies by these authors identified potential 12(S)-HETE binding sites in multiple subcellular compartments [186]. Specifically, ~ 66.6, 38, 18, and 24 fmol/20 x 106 cells of 12(S)-HETE binding sites are identified for cytosol, mitochondria, plasma membrane, and nuclei, respectively [186]. The cytosolic binding sites for 12(S)-HETE appear to be a huge macro- molecular aggregate of an approximate M.W. of 669,000 dalton which is sensitive to trypsin or chy- motrypsin [186]. Our own studies also suggest that 12(S)-HETE binds to specific (receptor) sites in tu- mor cells as well as in endothelial cells [194].

Based upon the above discussions, we present the following working hypothesis to explain the invol- vement of 12(S)-HETE in regulating tumor metas- tasis, as illustrated in Fig. 10 and 11. External 12(S)- HETE, through binding to the cell surface receptor, activates a pertussis toxin-sensitive G protein [193], which in turn activates PLC. Stimulation of PLC ac- tivates the phosphoinositide metabolism generat- ing from PIP 2 both DAG and IP3, which together activate PKC. Functionally activated PKC medi- ates protein (e.g., cytoskeletal elements) phospho- rylation resulting in a modulation of a variety of phenotypic properties of metastasizing tumor cells such as adhesion, spreading, and motility (Fig. 10). 13(S)-HODE antagonizes the 12(S)-HETE effects by binding to either the same binding site(s) for 12(S)-HETE or different binding site(s) which neg- atively affect the function of 12(S)-HETE receptors (Fig. 10). In another scheme (Fig. 11), tumor cell ad- hesion to endothelial cells activates the 12-1ipoxyge- nase, leading to arachidonic acid metabolism to 12(S)-HETE. 12(S)-HETE, by unknown mecha- nisms, activates PKC and translocates PKC to the

387

plasma membrane. PKC subsequently phosphory- lares and reorganizes the cytoskeletal elements which mobilize cdlb133 integrin-containing vesicles to the plasma membrane. Cytoskeletal rearrange- ments induced by 12(S)-HETE also are involved in modulation of tumor cell spreading on matrix and tumor cell motility. Tumor cell-derived 12(S)- HETE likewise acts on endothelial cells, resulting in enhanced ~v[33 integrin surface expression and tumor cell adhesion, and PKC-dependent phospho- rylation and reorganization of cytoskeleton and en- dothelial cell retraction. In addition, 12(S)-HETE may promote tumor cell release of cathepsin B to degrade the subendothelial matrix. In vivo, tumor cell interactions with platelets (through ~IIb[33 in- tegrins and many other receptors) will lead to plate- let aggregation, activation and release reactions. Platelet-derived 12(S)-HETE, which can be re- leased to the loci of tumor cell-platelet-endothelial cell interactions, can similarly act on endothelial cells to initiate the diverse biological responses de- scribed above, thus contributing to tumor cell me- tastasis (Fig. 1l).

Perspectives

Multiple determinants have been implicated in tu- mor progression and metastasis. No single factor alone has been known to determine the fate of me- tastasizing tumor cells. 12-Lipoxygenase obviously is among those factors that facilitate the metastatic process. This enzyme is found to be expressed in many types of tumor cells in addition to platelets, leukocytes, epidermal cells, and some other normal cells. We have obtained preliminary evidence that the platelet-type 12-1ipoxygenase is amplified at the mRNA level in certain human cancers and there ap- pears to be a positive correlation between the mRNA levels of 12-1ipoxygenase and the degree of malignancy [195]. Therefore, the possibility exists that 12-1ipoxygenase can be used clinically as a prognostic indicator for these cancers. Given the wide-spectrum modulatory role of 12(S)-HETE in tumor cell metastasis, functional 12-1ipoxygenase inhibitors such as baicalein and BHPP or their ana- logs may find applications in adjuvant therapies of

i ¸ ~

389

cancer patients aimed at interfering with or block- ing metastasis. Likewise, prostacylin and 13(S)- HODE, which are known to antagonize many of the 12(S)-HETE effects, may also be useful in can- cer treatment (see also chapter 11 for relevant dis- cussions). Furthermore, studies have demonstrated that 12(S)-HETE could represent a crucial interme- diate signaling molecule that mediates the func- tions of many growth factors and cytokines such as bFGF, PDGF, and AMF, etc. Thus, development of agents that directly interfere with or antagonize the functions of 12(S)-HETE may prove to be meaning- ful for the anti-signaling therapy of cancer metasta- sis, as currently advocated by many oncologists [187, 188].

Conclusion

- 12-Lipoxygenase catalyzes the transformation of AA to 12-HPETE which is then converted, either enzymatically or non-enzymatically, to 12(S)-HETE. There are at least three different types of mammalian 12-1ipoxygenases, which differ in their tissue distributions, preferential substrates, and primary products.

- The substrate specificity of 12-1ipoxygenases vs other lipoxygenases is determined by some well- conserved amino acid residues at or near the en- zyme active center. The presence of multiple up- stream activator sequences, in particular, an AP-2 site, in the promoter region of 12-1ipoxyge- nases provides the molecular basis for the induc- tion of the enzyme in various mammalian cells.

- Many solid tumor cells express 12-1ipoxygenase mRNA and protein. 12-Lipoxygenases ex- pressed in tumor cells are regulated by multiple extra- and intracellular factors. The ability of tu- mor cells to biosynthesize 12(S)-HETE is a posi- tive correlate of their metastatic potential.

- 12(S)-HETE is a key intracellular signaling mol-

ecule that activates protein kinase C. Exogenous and tumor cell-derived 12(S)-HETE modulates a wide variety of properties of tumor cells (ad- hesion, spreading, invasion, motility, etc) and en- dothelial cells (retraction, proliferation, adhe- sion and migration, etc). 12-Lipoxygenase and 12(S)-HETE may repre- sent key molecules to which anti-cancer and an- ti-metastatis therapeutic regimens are targeted.

Key unanswered questions

- How many different types of 12-1ipoxygenases are expressed by a distinct type of tumor? Are there any key differences between tumor cell ex- pressed 12-1ipoxygenases and their counterparts in normal cells in terms of substrate specificity and profile of the metabolites?

- How does 12(S)-HETE exert its biological func- tions in tumor cells and endothelial cells? Are there authentic 12(S)-HETE binding sites or re- ceptors? How does 12(S)-HETE activate pro- tein kinase C?

- How can we exploit our knowledge about 12-1i- poxygenase and 12(S)-HETE to develop effica- cious anti-cancer and anti-metastasis protocols?

Acknowledgements

Original work in the authors' laboratories was sup- ported by National Institute of Health grants CA47115, CA 29997 (KVH), NATO Linkage Grant Programs LG92311 and the Fogarty International Research grant (KVH and JT) and grants from the Development Office of Harper Hospital and Wayne State University (KVH).

<

Fig. 1l. Role of 12-Iipoxygenase and 12(S)-HETE in tumor cell-vasculature interactions. Details are given in the text. TC, tumor cell; EC, endothelial cells, SEM, subendothelial matrix; 12-LOX, 12-1ipoxygenase; AA, arachidonic acid; CSK, cytoskeleton; PKC, protein kinase C; PGI2, prostacyclin. Modified after 98.

390

References

1. Weiss L, Orr FW, Honn KV: Interactions between cancer and the microvasculature: a rate-regulator for metastasis. Clin Expl Metastasis 7: 127-167, 1989

2. Liotta LA: Tumor invasion and metastasis: Role of the ex- tracellular matrix. Cancer Res 46: 1-7, 1986

3. Zetter BR: The cellular basis of site-specific tumor metas- tasis. N Engl J Med 322: 605-612,1990

4. Nicolson GL: Tumor and host molecules important in the organ preference of metastasis. Seminars Cancer Biol 2: 143-154,1991

5. Weinstat-Saslow D, Steeg PS: Angiogenesis and coloniza- tion in the tumor metastatic process: basic and applied ad- vances. FASEB J 8: 401-407, 1994

6. Needleman P, Turk J, Jakschik BA, Morrison AR, Lefko- with JB: Arachidonic acid metabolism. Ann Rev Biochem 55: 69-102, 1986

7. Spector AA, Gordon JA, Moore SA: Hydroxyeicosate- traenoic acids (HETEs). Prog Lipid Res 27: 271-273, 1988

8. Serhan CN: Lipoxin biosynthesis and its impact in inflam- matory and vascular events. Biochim Biophys Acta 1212: 1-25, 1994

9. Nicosia S, Patrono C: Eicosanoid biosynthesis and action: novel opportunities for pharmacological intervention. FA- SEB J 3: 1941-1948, 1989

10. Greenberg ER, Baron JA, Freeman DH, Jr., Mandel JS, Haile R: Reduced risk of large-bowel adenomas among as- pirin users. J Natl Cancer Inst 85: 912-916, 1993

11. Felder CC, Ma LL, Liotta LA, Kohn EC: The antiprolifer- ative and antimetastatic compound L651582 inhibits mus- carinic acetylcholine receptor-stimulated calcium influx and arachidonic acid release. J Pharmacol Expl Ther 267: 967-971, 1991

12. Marnett LJ: Aspirin and the potential role of prostaglan- dins in colon cancer. Cancer Res 52: 5575-5589,1992

13. Butcher RD, Wojcik SJ, Lints T, Wilson T, Schofleld PC, Ralph R: Arachidonic acid, a growth signal in murine P815 mastocytoma cells. Cancer Res 53: 3405-3410, 1993

14. Honn KV, Cicone B, Skoff A: Prostacyclin, A potent anti- metastatic agent. Science 212: 1270-1272,1981

15. H°nn KV'Tang DG' Chen YQ: Platelets and cancer metas- tasis: More than an epiphenomenon. Seminar Thromb He- most 18: 392-415, 1993

16. Honn KV, Tang DG, Crissman JD: Platelets and cancer metastasis: A causal relationship? Cancer Metastasis Rev 11: 325-351,1993

17. Chen YQ, Liu B, Tang DG, Honn KV: Fatty acid modula- tion of tumor cell-platelet-vessel wall interaction. Cancer Metastasis Rev 11: 389-410,1993

18. Honn KV, Tang DG: Adhesion molecules and tumor cell interaction with endothelium and subendothelial matrix. Cancer Metastasis Rev 11: 353-375, 1992

19. Tang DG, Honn KV: Eicosanoids and tumor cell metasta- sis. In: Harris JE, Braun DR Anderson KM (eds) Prosta- glandin Inhibitors in Tumor Immunology and Immunoth-

erapy. CRC press, Boca Raton, Ann Arbor, London, To- kyo, 1994, pp 73-108

20. Bastida E, Bertomeu MC, Haas TA, Almiral L, Lauri D, Orr FW, Buchanan MR: Regulation of tumor cell adhesion by intracellular 13-HODE:15HETE ratio. J Lipid Med 2: 281-293, 1990

21. Buchanan MR, Bastida E: Endothelium and underlying membrane reactivity with platelets, leukocytes and tumor cells: Regulation by the lipoxygenase-derived fatty acid metabolites, 13-HODE and HETE's. Med Hypothesis 27: 317-325, 1988

22. Funk CD: Molecular biology in the eicosanoid field. Prog Nuc Acid Res Mol Bio145: 67-98, 1993

23. Takahashi Y, Ueda N, Yamamoto S: Two immunologically and catalytically distinct arachidonate 12-Lipoxygenases of bovine platelets and leukocytes. Arch Biochem Biophys 266: 613-621, 1988

24. Hansbrough JR, Takahashi Y, Ueda N, Yamamoto S, Holtzman MJ: Identification of a novel arachidonate 12- lipoxygenase in bovine tracheal epithelial cells distinct from leukocyte and platelet forms of the enzyme. J Biol Chem 265: 1771-1776, 1990

25. Yoshimoto T, Suzuki H, Yamamoto S, Takai T, Yokoyama C, Tanabe T: Cloning and sequence analysis of the cDNA for arachidonate 12-1ipoxygenase of porcine leukocytes. Proc Natl Acad Sci USA 87: 2142-2146, 1990

26. De Marzo N, Sloane DL, Dicharry S, Highland E, Sigal E: Cloning and expression of an airway epithelial 12-1ipoxyge- nase. Am J Physiol 262: L198-L207, 1992

27. Sigal E, Craik CS, Highland E, Grunberger D, Costello LL, Dixon RAF, Nadel JA: Molecular cloning and primary structure of human 15-1ipoxygenase. Biochem Biophys Res Commun 157: 457.464, 1988

28. O'Prey J, Chester J, Thiele BJ, Janetzki S, Prehn S, Fleming J, Harrison PR: Promoter structure and complete se- quence of the gene encoding the rabbit erythroid cell-spe- cific 15-1ipoxygenase. Gene 84: 493-499, 1989

29. Dixon RAF, Jones RE, Diehl RE, Bennett CD, Kargman S, Rouzer CA: Cloning of the cDNA for human 5-1ipoxyge- nase. Pro Natl Acad Sci USA 85: 416-420, 1988

30. Balcarek JM, Theisen TW, Cook MN, Varrichio A, Hwang S, Strohsacker MW, Crooke ST: Isolation and character- ization of a cDNA clone encoding rat 5-1ipoxygenase. J Biol Chem 263: 13937-13941, 1988

31. Funk CD, Furci L, Fitzgerald GA: Molecular cloning, pri- mary structure, and expression of the human platelet/ erythroleukemia cell 12-lipoxygenase. Proc Natl Acad Sci USA 87: 5638-5642, 1990

32. Watanabe S, Medina JE Haeggstrom JZ, Radmark O, Sa- muelsson B: Molecular cloning of a 12-1ipoxygenase cDNA from rat brain. Eur J Biochem 212: 605-612,1992

33. Izumi T, Hoshiko S, Radmark O, Samuelsson B: Cloning of the cDNA for human 12-1ipoxygenase. Proc Natl Acad Sci USA 87: 7477-7481, 1990

34. Sloane DL, Craik SS, Sigal E: The expression of active hu-

man reticulocyte 15-1ipoxygenase in bacteria. Biochim Bio- phys Acta 49: $11-$16, 1990

35. Funk CD, Funk LB, FitzGerald GA, Samuelsson B: Char- acterization of human 12-1ipoxygenase genes. Proc Natl Acad Sci USA 89: 3962-3966, 1992

36. Funk CD, Furchi L, Fitzgerald GA: Adv Prostaglandin Thromboxane Leukotriene Res 21: 33, 1990

37. Hada T, Ueda N, Takahashi Y, Yamamoto S: Catalytic properties of human platelet 12-1ipoxygenase as compared with the enzymes of other origins. Biochim Biophys Acta 1083: 89, 1991

38. Kuhn H, Sprecher H, Brash AR: On singular and dual posi- tional specificity of lipoxygenases. J Biol Chem 265:16300- 16306, 1990

39. Boyington JC, Gaffney B J, Amzel GLM: The three-di- mensional structure of an arachidonic acid 15-1ipoxyge- nase. Science 260: 1482--1486, 1993

40. Chen X-S, Kurre U, Jenkins NA, Copeland NG, Funk CD: cDNA cloning, expression, mutagenesis of C-terminal iso- leucine, genomic structure, and chromosomal localization of murein 12-1ipoxygenases. J Biol Chem 269:13979-13987, 1994

41. Fleming J, Thiele B J, Chester J, O'Prey J, Janetzki S, Air- ken A, Anton IA, Rapoport SM, Harrison PR: The com- plete sequence of the rabbit erythroid cell-specific 15-1i- poxygenase mRNA: comparison of the predicted amino acid sequence of the erythrocyte lipoxygenase with other lipoxygenases. Gene 79: 18-188, 1989

42. Sloane DL, Leung R, Craik CS, Sigal E: A primary deter- minate for lipoxygenase positional specificity. Nature 354: 149-152, 1991

43. Matsumoto T, Funk CD, Radmark O, Hoog JO, Jornavall H, Samuelsson B: Molecular cloning and amino acid se- quence of human 5-1ipoxygenase. Proc natl Acad Sci USA 85: 26-30, 1988

44. Nguyen T, Falgueyret J-R Abramovitz M, Riendeau D: Evaluation of the role of conserved His and Met residues among lipoxygenases by site-directed mutagenesis of re- combinate human 5-1ipoxygenase. J Biol Chem 266: 22057-22062, 1991

45. Zhang YY, Radmark O, Samuelson B: Mutagenesis of some conserved residues in human 5-1ipoxygenase: Effects on enzyme activity. Proc Natl Acad Sci USA 89: 485-489, 1992

46. Yoshimoto T, Arakaws T, Hada T, Yamamoto S, Takahashi E: Structure and chromosomal localization of human ara- chidonate 12-1ipoxygenase gene. J Biol Chem 267: 24805- 14809, 1992

47. Arakawa T, Oshima T, Kishimoto K, Yoshimoto T, Yama- moto S: Molecular structure and function of the porcine arachidonate 12-lipoxygenase gene. J Biol Chem 267: 12188-12191, 1992

48. Chang W-C, Liu Y-W, Ning C-C, Suzuki H, Yoshimoto T, Yamamoto S: Induction of Arachidonate 12qipoxygenase mRNA by epidermal growth factor in A431 cells. J Biol Chem 268: 18734-18739, 1993

391

49. Chang W-C, Ning C-C, Lin MT, Huang J-D: Epidermal growth factor enhances a microsomal 12-lipoxygenase ac- tivity in A431 cells. J Biol Chem 267: 3657-3666, 1992

50. Funk CD, Fitzgerald GA: Eicosanoid forming enzyme mRNA in human tissues. Analysis by quantitative polym- erase chain reaction. J Biol Chem 266: 12508-12513, 1991

51. Nataraj an R, Gu J-L, Rossi J, Gonzales N, Lanting L, Xu L, Nadler J: Elevated glucose and angiotensin II increase 12- lipoxygenase activity and expression in porcine aortic smooth muscle cells. Proc Natl Acad Sci USA 90: 4947- 4951, 1993

52. Faisst S, Meyer S: Compilation of vertebrate-encoded tran- scription factors. Nucl Acids Res 20: 3-26, 1992

53. Chen YQ, Duniec ZM, Liu B, Hagmann W, Gao X, Shimoji K, Marnett LJ, Johnson CR, Honn KV: Endogenous 12(S)- HETE production by tumor cells and its role in metastasis. Cancer Res 54: 1574-1579, 1994

54. Glasgow WC, Eling TE: Epidermal growth factor regula- tion of linoleic acid metabolism in syrian hamster embryo fibroblasts. In: Nigam S, Honn KV, Marnett L J, Walden T (eds) Eicosnois and other Bioactive Lipids in Cancer In- flammation and Radiation Injury. Kluwer Academic Pub- lishers, 1992, pp 467-470

55. Hussain H, Shornick JR Shannon VR, Wilson JD, Funk CD, Pentland AR Holtzman M J: Epidermis contains plate- let-type 12-1ipoxygenase that is overexpressed in germinal layer keratinocytes in psoriasis. Am J Physiol 266: C243- C253, 1994

56. Chen X-S, Funk CD: Structure-function properties of hu- man platelet 12-1ipoxygenase: Chimeric enzyme and in vit-

ro mutagenesis studies. FASEB J 7: 694-701, 1993 57. Kuhn H, Heydeck D, Sprecher H: On the mechanistic rea-

sons for the dual positional specificity of the reticulocyte lipoxygenase. Biochim Biophys Acta 830: 25-29, 1985

58. Gardner HW: Soybean lipoxygenase-1 enzymically forms both (9S)- and (13S)-hydroperoxides from linoleic acid by pH-dependent mechanism. Biochim Biophys Acta 1001: 274-281, 1989

59. Kuhn H, Heydeck D, Wiesner R, Schewe T: The positional specificity of wheat lipoxygenase. The carboxylic group as signal for the recognition of the hydrogen removal. Bio- chim Biophys Acta 830: 25-29, 1985

60. Rapoport S, Hartel B, Hausdorf G: Methionine sulfoxide formation: the cause of self-inactivation of reticulocyte li- poxygenase. Eur J Biochem 139: 573-576, I984

61. Zakut R, Grossman S, Pinsky A, Wilchek M: Evidence for an essential methionine residue in lipoxygenase. FEBS Letters 71: 107-110, 1976

62. Watanabe T, Haeggstrom JZ: Rat 12-1ipoxygenase: muta- tions of amino acids implicated in positional specificity of 15- and 12-1ipoxygenases. Biochem Biophys Res Commun 192: 1023-1029,1993

63. Yamamoto S: Mammalian lipoxygenases: molecular struc- tures and functions. Biochim Biophys Acta 1128: 117-131, 1992

64. Toh H, Yokoyama C, Tanabe T, Yoshimoto T, Yamamoto

392

S: Molecular evolution of cyclooxygenase and lipoxyge- nase. Prostaglandins 44: 291-315, 1992

65. Sigal E: The molecular biology of mammalian arachidonic acid metabolism. Amer J Physiol 4(1): L13-L28, 1991

66. Kroneck PMH, Cucurou C, Ullrich V, Ueda N, Suzuki H, Yoshimoto T, Matsuda S, Yamamoto S: Porcine leukocyte 5- and 12-1ipoxygenases are iron enzymes. FEBS Letters 286: 105-107,1991

67. Schilstra MJ, Veldink GA, Vliegenthart JFG: Kinetic anal- ysis of the induction period in lipoxygenase catalysis. Bio- chem 32: 7686-7691, 1993

68. Aoshima H, Kajiwara T, Hatanaka A, Nakatani H, Hiromi K: Modification of lipoxygenase by hydrogen peroxide and photooxidation. Int J Pept Protein Res 10: 219-225,1977

69. Butovich IA: Aerobic catalytic cycle of 5-1ipoxygenase: a new interpretation of the enzyme activation by synthetic and natural amphiphilic acids. In: The Proceedings for the Third International Conference on Lipid Mediators in Health and Disease. Jerusalem, Israel. 1993, pp 121

70. Feiters MC, Boelens H, Veldink GA, Vliegenthart JFG, Navaratnam S, Allen JC, Nolting H-F, Hermes C: X-Ray absorption spectroscopic studies on iron in soybean lipoxy- genase: a model for mammalian lipoxygenases. Recl Trav Chim Pays Bas 109: 133-146, 1990

71. Nelson MJ, Seitz SR Cowling RA: Enzyme-bound penta- dienyl and peroxyl radicals in purple lipoxygenase. Bio- chem 29: 6897-6903, 1990 Veldink GA, Vliegenthart JFG, Boldingh G: Plant lipoxy- genases. Prog Fat Other Lipids 15: 131-166, 1977 Bryant RW, Schewe T, Rapoport SM, Bailey JM: Leuko- triene formation by a purified reticulocyte lipoxygenase enzyme. Conversion of arachidonic acid and 15-hydrope- roxyeicosatetraenoic acid to 14,15-1eukotriene A 4. J Biol Chem 260: 3548-3555, 1985 Cucurou C, Battioni JR Daniel R, Mansuy D: Peroxidase- like activity of lipoxygenases: different substrate specificity of potato-5 lipoxygenase and soybean 15-1ipoxygenase, and particular activity of vitamine E derivatives for the 5- lipoxygenase. Biochim Biophys Acta 1081:95 105, 1991 Kuhn H, Eggert L, Zabolotsky OA, Myagkova G, Schewe T: Keto fatty acids not containing doubly allylic methylenes and lipoxygenase substrates. Biochemistry 30: 1026% 10273, 1991 Funk MO, Jr., Andre JC, Otsuki T: Oxygenation of trans polyunsaturated fatty acids by lipoxygenase reveals steric features of the catalytic mechanism. Biochem 26: 6880- 6889, 1987

77. Salzmann U, Kuhn H, Schewe T, Rapoport SM: Pentane formation during the anaerobic reaction of reticulocyte li- poxygenase. Comparison with lipoxygenases from soy- beans and green pea seeds. Biochim Biophys Acta 795: 535-542,1984

78. Takahashi Y, Reddy GR, Ueda N, Yamamoto S, Arase S: Arachidonate 12-1ipoxygenase of platelet-type in human epidermal cells. J Biol Chem 268:16443-16448,1993

79. Mahmud I, Suzuki T, Yamamoto Y, Suzuki H, Takahashi 5I,

72.

73.

74.

75.

76.

Yoshimoto T, Yamamoto S: Induction of cyclooxygenase and suppression of 12-1ipoxygenase in human erythroleu- kemia cells upon phorbol ester-induced differentiation. Biochim Biophys Acta 1166: 211-216, 1993

80. Siegel MI, McConnell RT, Porter NA, Cuatrecasas P: Ara- chidonate metabolism via lipoxygenase and 12L-hydrope- roxy-5, 8, 10, 14-icosatetraenoic acid peroxidase sensitive to anti-inflammatory drugs. Proc Natl Acad Sci USA 77: 308-312, 1980

81. Yoshimoto T, Miyamoto Y, Ochi K, Yamamoto S: Arachid- onate 12-1ipoxygenase of porcine leukocyte with activity for 5-hydroxyeicosatetraenoic acid. Biochem Biophys Ac- ta 713: 638-646,1982

82. Marnett LJ, Leithauser MT, Richards KM, Blair I, Honn KV, Yamamoto S, Yoshimoto T: Arachidonic acid metabo- lism of cytosolic fractions of Lewis Lung carcinoma cells. Adv Prostaglandin Thromboxane Leukotriene Res 21: 895-900, 1990

83. Shornick LR Holtzman M J: A cryptic, microsomal-type arachidonate 12-1ipoxygenase is tonically inactivated by oxidation-reduction conditions in cultured epithelial cells. J Biol Chem 268: 371-376, 1993

84. Hagmann W, Kagawa D, Renaud C, Honn KV: Activity and protein distribution of 12-1ipoxygenase in HEL cells: Induction of membrane-association by phorbol ester TPA, modulation of activity by glutathione and 13-HPODE, and Ca2+-dependent translocation to membranes. Prostaglan- dins 46: 471-477, 1993 .

85. Hagmann W, Maher R, Honn KV: Intracellular distribu- tion, activity, and Ca2+-dependent translocation of 12-1i- poxygenase in Lewis lung tumor cells. In: Eicosanoids and Other Bioactive Lipids in Cancer, Inflammation, and Ra- diation Injury (K.V. Honn, L.J. Marnett, S. Nigam, eds), Kluwer Academic Press, in Press, 1994

86. Baba A, Sakuma S, Okamoto H, Inoue T, Iwata H: Calci- um induces membrane translocation of 12-1ipoxygenase in rat platelets. J Biol Chem 264:15790-15795,1989

87. Liu B, Timar J, Howlett J, Diglio CA, Honn KV: Lipoxyge- nase metabolites of arachidonic and linoleic acids modu- late the adhesion of tumor cells to endothelium via regu- lation of protein kinase C. Cell Regul 2: 1045-1055, 1991

88. Schonhardt T, Ferber E: Translocation of phospholipase A2 from cytosol to membranes induced by 1-oleoyl-2-ace- tyl-glycerol in serum-free cultured macrophages. Biochem Biophys Res Commun 149: 769-775, 1987

89. Balsinde J, Diez E, Mollinedo F: Arachidonic acid release from diacylglycerol in human neutrophils. J Biol Chem 266: 15638-15643,1991

90. Rouzer CA, Kargman S: Translocation of 5-1ipoxygenase to the membrane in human leukocytes challenged with ionophore A23187. J Biol Chem 263: 10980-10988,1988

91. Furstenberger G, Hagedorn H, Jacobi T, Besemfelder E, Stephan M, Lehmann WD, Marks F: Characterization of an 8-1ipoxygenase activity induced by the phorbol ester tu- mor promoter 12-O-tetradecanoylphorbol-13-acetate in mouse skin in vivo. J Biol Chem 266: 15738-15745, 1991

92. Clark JD, Lin LL, Kriz RW, Ramesha CS, Sultzman LA, Lin AY, Milona N, Knopf JL: A novel arachidonic acid- selective cytosolic PLA 2 contains a Ca2+-dependent trans- location domain with homology to PKC and GAR Cell 65: 1043-1051, 1991

93. Hadjiagapiou C, Spector AA: 12-hydroxyeicosatetraenoic acid reduces prostacyclin production by endothelial cells. Prostaglandins 31: 1135-1144, 1986

94. Sekiya F, Takagi J, Usui T, Kawaiiri K, Kobayashi Y, Sato E Saito Y: 12S-hydroxyeicosatetraenoic acid plays a central role in the regulation of platelet activation. Biochem Bio- phys Res Commun 179: 345-351, 1991

95. Hofer G, Bieglmayer CH, Kopp B, Janisch H: Measure- ment of eicosanoids in menstrual fluid by the combined use of high pressure chromatography and radioimmunoassay. Prostaglandins 45: 413-426, 1993

96. Wetzka B, Schafer W, Scheibel M, Nusing R, Zahradnik HP: Eicosanoid production by intrauterine tissues before and after labor in short-term tissue culture. Prostaglandins 45: 571-581, 1993

97. Liu B, Marnett LJ, Chaudhary A, Ji C, Blair IA, Johnson CR, Diglio CA, Honn KV: Biosynthesis of 12(S)-hydrox- yeicosatetraenoic acid by BI6 amelanotic melanoma cells is a determinant of their metastatic potential. Lab Invest 70: 314-323, 1994

98. Honn KV, Tang D, Grossi I, Duniec ZM, Timar J, Renaud C, Leithauser M, Blair 1, Johnson CR, Diglio CA, Kimler VA, Taylor JD, Marnett L J: Tumor cell-derived 12(S)-Hy- droxyeicosatetraenoic acid induces microvascular endo- thelial cell retraction. Cancer Res 54: 565-574, 1994

99. Honn KV, Nelson KK, Renaud C, Bazaz R, Diglio CA, Timar J: Fatty acid modulation of tumor cell adhesion to microvessel endothelium and experimental metastasis. Prostaglandins 44: 413-429, 1992

100. Honn KV, Tang DG, Grossi 1M, Renaud C, Duniec ZM, Johnson CR, Diglio CA: Enhanced endothelial cell retrac- tion mediated by 12(S)-HETE: A proposed mechanism for the role of platelets in tumor cell metastasis. Exp Cell Res 210: 1-9, 1994

101. Pauli BU, Augustin-Voss HG, EI-Sabban ME, Johnson RC, Hammar DA: Organ preference of metastasis: the role of endothelial cell adhesion molecules. Cancer Metastasis Rev 9: 175-189, 1990

102. Ruoslahti E, Giancotti FG: Integrins and tumor cell dis- semination. Cancer Cells 4: 119-126, 1989

103. Singer I J, Scott S, Kawka DW, Kazazis DM: Adhesosomes: specific granules containing receptors for laminin, c3bi, fi- brinogen, fibronectin and vitronectin in human polymor- phonuclear leukocytes and monocytes. J Cell Biol 109: 3169-3182, 1989

104. Chopra H, Timar J, Chen YQ, Rong XH, Grossi IM, Fitz- gerald LA, Taylor JD, Honn KV: The lipoxygenase metab- olite 12(S)-HETE induces a cytoskeleton-dependent in- crease in surface expression of integrin cdlb]33 on melan- oma cells. Int J Cancer 49: 774-786, 1991

105. Zachary I, Rozengurt E: Focal adhesion kinase (p125FA~):

393

A point of convergence in the action of neuropeptides, in- tegrins, and oncogenes. Cell 71: 891-894, 1992

106. Schaller MD, Borgman CA, Cobb BS, Vines RR, Reynolds AB, Parsons TJ: pp125 FAK, a structurally distinctive pro- tein-tyrosine kinase associated with focal adhesions. Proc Natl Acad Sci USA 89: 5192-5196, 1992

107. Guan J-L, Shalloway D: Regulation of focal adhesion-as- sociated protein tyrosine kinase by both cellular adhesion and oncogenic transformation. Nature 358: 690-692, 1992

108. Grossi IM, Fitzgerald LA, Umbarger LA, Nelson KK, Di- glio CA, Taylor JD, Honn KV: Bidirectional control of membrane expression and/or activation of the tumor cell IRGplIb/l lIa receptor and tumor cell adhesion by lipoxy- genase products of arachidonic and linoleic acid. Cancer Res 49: 1029-1037, 1989

109. Chang YS, Chen YQ, Timar J, Grossi IM, Fitzgerald LA, Diglio CA, Honn KV: Increased expression of 0dl13b3 in- tegrin in subpopulations of murine melanoma cells with high lung-colonizing ability. Int J Cancer 51: 445-454, 1992

110. Chen YQ, Gao X, Timar J, Tang D, Grossi IM, Chelladurai M, Kunicki TJ, Fligiel SEG, Taylor JD, Honn KV: Identifi- cation of the odlbl33 integrin in murine tumor cells. J Biol Chem 267: 17314-17320, 1992

111. Chopra H, Timar J, Rong X, Grossi IM, Hatfield JS, Fligiel SEG, Finch CA, ~I:aylor JD, Honn KV: Is there a role for the tumor cell integrin cdl[33 and cytoskeleton in tumor cell- platelet interaction? Clin Expl Metastasis 10: 125-138,1992

112. Timar J, Chen YQ, Liu B, Bazaz R, Taylor JD, Honn KV: The lipoxygenase metabolite 12(S)-HETE promotes aIfb[33 integrin-mediated tumor cell spreading on fibro- nectin. Int J Cancer 52: 594-603, 1992

113. Stoker M, Gherardi E: Regulation of cell movement: the motility cytokines. Biochim Biophys Acta 1072: 81-102, 1991

114. Liotta LA. Mandler R, Murano G, Katz DA, Gordon RK, Chiang PK, Schiffman E: Tumor-cell autocrine motility factor. Proc Natl Acad Sci USA 83: 3302-3306, 1986

115. Watanabe H, Carmi R Hogan V, Raz T, Silletti S, Nabi IR, Raz A: Purification of human tumor cell autocrine motility factor and molecular cloning of its receptor. J Biol Chem 266: 13442-13448, 1991

116. Timar J, Silletti S, Bazaz R, Raz A, Honn KV: Regulation of melanoma-cell motility by the lipoxygenase metabolite I2(S)-HETE. Int J Cancer 55: 1003-1010, 1993

117. Silletti S, Raz A: Autocrine motility factor (AMF) is a growth factor. Biochem Biophys Res Commun 194: 446- 457, 1993

118. Liu B, Maher R J, Hannun YA, Porter AT, Honn KV: 12(S)- HETE increases the invasive potential of prostate tumor cells through selective activation of PKCa. J Natl Cancer Inst 86: 1145-1151, t994

119. Honn KV, Timar J, Rozhin J, Bazaz R, Sameni M, Ziegler G, Sloane BF: A lipoxygenase metabolite, 12-(S)-HETE, stimulates protein kinase C-mediated release of cathepsin B from malignant cells. Exp Cell Res 214: 120-130, 1994

394

120. Schmitt M, Graft JH: Tumor-associated proteases. Fibrino- lysis 6: 3-26, 1992

121. Sloane BF, Moin K, Krepela E, Rozhin J: Cathepsin B and its endogenous inhibitors: the role in tumor cell malignan- cy. Cancer Metastasis Rev 9: 333-352, 1990

122. Sloane BF, Moin K, Sameni M, Tait LR, Rozhin J, Ziegler J: Membrane association of cathepsin B can be induced by transfection of human breast epithelial cells with c-Ha-ras oncogene. J Cell Sci 107: 373-384, 1994

123. Sloane BF, Rozhin J, Johnson K, Taylor H, Crissman JD, Honn KV: Cathepsin B: association with plasma mem- brane in metastatic tumors. Proc Natl Acad Sci USA 83: 2483-2487, 1986

124. Sloane BF, Rozhin JR, Gomez AR Grossi IM, Honn KV: Effects of 12-hydroxyeicosatetraenoic acid on release of cathepsin B and cysteine proteinase inhibitors from malig- nant melanoma cells. In: Honn KV, Marnett L J, Nigam S, Walden TL (eds) Eicosanoids and other Bioactive Lipids in Cancer and Radiation Injury. Kluwer, Boston, 1991, pp 373-377

125. Timar J, Tang D, Bazaz R, Haddard MM, Kimler VA, Tay- lor JD, Honn KV: PKC mediates 12(S)-HETE-induced cy- toskeletal rearrangement in B16a melanoma cells. Cell Motil Cytoske126: 49-65, 1993

126. Tang DG, Honn KV: Role of protein kinase C and phos- phatases in 12(S)-HETE-induced tumor cell cytoskeletal reorganization. In: Honn KV, Marnett LJ, Nigam S (eds) Eicosanoids and Other Bioactive Lipids in Cancer, Inflam- mation and Radiation Injury. Kluwer, Boston, 1994, in press

127. Crissman JD, Hatfield J, Schadenbrand M, Sloane BF, Honn KV: Arrest and extravasation of B16 amelanotic melanoma in murine lungs. A light and electron micro- scopic study. Lab Invest 53: 470-478, 1985

128. Chew EC, Cheung SL: In vitro studies on the invasion of blood vessel by choriocarcinoma cells. Anticancer Res 9: 133-139, 1989

129. Kramer RH, Nicolson GL: Interactions of tumor cells with vascular endothelial cell monolayers: a model for metastat- ic invasion. Proc Natl Acad Sci USA 76: 5704-5708, 1979

130. Huang A J, Silverstein SC: Mechanisms of neutrophil mi- gration across endothelial cells. In: Simionescue N, Simino- nescu M (eds) Endothelial Cell Dysfunctions. Plenum Press, New York, 1992, pp 201-231

131. Lafrenie R, Shaughnessy S, Orr FW: Cancer cell interac- tions with injured or activated endothelium. Cancer Me- tastasis Rev 11: 377-388, 1992

132. Nicholson GL, Custead SE, Dulski KM, Milas L: Effect of gamma irradiation on cultured rat and mouse microvessel endothelial cells: metastatic tumor cell adhesion, subendo- thelial matrix degradation, and secretion of tumor cell growth factors. Clin Exp Metastasis 9: 457-468,1991

133. Honn KV, Grossi IM, Diglio CA, Wotjukiewicz M, Taylor JD: Enhanced tumor cell adhesion to the subendothelial matrix resulting from 12(S)-HETE-induced endothelial cell retraction. FASEB J 3: 2285-2293, 1989

134. Grossi IM, Diglio CA, Honn KV: Control of tumor cell in- duced endothelial cell retraction by lipoxygenase metabo- lites, prostacyclin and prostacyclin analogs. In: Honn KV, Marnett LJ, Nigam S, Walden T (eds) Eicosanoids and Other Bioactive Lipids in Cancer and Radiation Injury. Kluwer Academic Publishers, Boston, 1991, pp 409--414

135. Tang DG, Diglio CA, Honn KV: 12(S)-HETE-induced mi- crovascular endothelial cell retraction results from PKC- dependent rearrangement of cytoskeletal elements and c~v[53 integrins. Prostaglandins 45: 249-268, 1993

136. Tang DG, Chen YQ, Diglio CA, Honn KV: PKC-depend- ent effects of 12(S)-HETE on endothelial cell vitronectin receptor and fibronectin receptor. J Cell Bio1121: 689-704, 1993

137. Tang DG, Timar J, Grossi IM, Renaud C, Kimler VA, Di- glio CA, Taylor JD, Honn KV: The lipoxygenase metabo- lite, 12(S)-HETE, induces a protein kinase C-dependent cytoskeletal rearrangement and retraction of microvascu- lar endothelial cells. Exp Cell Res 207: 361-375, 1993

138. E1-Sabban M, Pauli BU: Cytoplasmic dye transfer between metastatic tumor cells and vascular endothelium. J Cell Bio1115: 1375-1382,1991

139. Wysolmersky RB, Lagunoff D: Inhibition of endothelial cell retraction by ATP depletion. Am J Patho1132: 28-37, 1988

140. Laposata M, Dovnasky DK, Shen HS: Thrombin-induced gap formation in confluent endothelial cell monolayers. Blood 62: 549-556, 1983

141. Wysolmersky RB, Lagunoff D: Involvement of myosin light chain kinase in endothelial cell retraction. Proc Natl Acad Sci USA 87: 16-20, 1990

142. Tang DG, Diglio CA, Honn KV: 12(S)-HETE-induced mi- crovascular endothelial cell retraction is mediated by cy- toskeletal rearrangement dependent on PKC activation. In: Nigam S, Honn KV, Marnett L J, Walden T (eds) Eicosa- holds and Other Bioactive Lipids in Cancer, Inflammation and Radiation Injury. Kluwer Academic Publishers, Bos- ton, 1992, pp 219-229

143. Albelda SM, Mueller WA, Buck CA, Newman PJ: Mole- cular and cellular properties of PECAM-1 (endoCAM/ CD31): A novel vascular cell-cell adhesion molecule. J Cell Bio1114: 1059-1068, 1991

144. Newman PJ, Berndt MC, Gorski J, White GC, Lyman S, Paddock C, Mueller WA: PECAM-1 (CD31) cloning and relation to adhesion molecules of the immunoglobulin gene superfamily. Science 247: 1219-1222,1990

145. Tang DG, Chen YQ, Newman P J, Shi L, Gao X, Diglio CA, Honn KV: Identification of PECAM-1 in solid tumor cells and its potential involvement in tumor cell adhesion to en- dothelium. J Biol Chem 268: 22883-22894, 1993

146. Tang DG, Grossi IM, Chert YQ, Diglio CA, Honn KV: 12(S)-HETE promotes tumor cell adhesion by increasing surface expression of ctv~3 integrins on endothelial cells. Int J Cancer 54: 102-111, 1993

147. Tang DG, Diglio CA, Honn KV: Activation of microvascu- lar endothelium by eicosanoid 12(S)-hydroxyeicosatetrae-

noic acid leads to enhanced tumor cell adhesion via up-reg- ulation of surface expression of av[33 integrin: A posttran- scriptional, protein kinase C- and cytoskeleton-dependent process. Cancer Res 54: 1119-1129, 1994

148. Hamberg M, Samuelsson B: Prostaglandin endoperoxides. Novel transformation of arachidonic acid in human plate- lets. Proc Natl Acad Sci USA 71: 3400-3404, 1974

149. Setty BNY, Dubowy RL, Stuart M J: Endothelial cell pro- liferation may be mediated via the production of endoge- nous lipoxygenase metabolites. Biochem Biophys Res Commun 144: 345-351, 1987

150. Axelrod J, Butch RM, Jelsema C: Receptor-mediated acti- vation of phospholipase A z via GTP-binding proteins: ar- chidonic acid and its metabolite as second messengers. TINS 11: 106-113, 1988

151. Perlman MB, Johnson A, Jubiz W. Malik AB: Lipoxyge- nase products induce neutrophil activation and increase endothelial permeability after thrombin-induced pulmo- nary microembolism. Circulation Res 64: 62-73, 1989

152. Rozengurt E: A role of arachidonic acid and its metabo- lites in the regulation of p21 ras activity. Cancer Cells 3: 397- 398, 1991

153. Hannigan GE, Williams BRG: Signal transduction by in- terferon-c~ through arachidonic acid metabolism. Science 251: 204-207,1991

154. Smith R J, Justen JM, Nidy EG, Sam LM, Bleasdale JE: Transmembrane signaling in human polymorphonuclear neutrophils: 15(S)-hydroxy-(5Z, 8Z, l lZ, 13E)-eicosate- traenoic acid modulates receptor agonist-triggered cell ac- tivation. Proc Natl Acad Sci USA 90: 7270-7274, 1993

155. Han JW, McCormick F, Macara IG: Regulation of Ras- GAP and the neurofibromatosis-1 gene product by eicosa- noids. Science 252: 576-579, 1991

156. Valles J, Santos MT, Marcus A J, Safier JB, Broekman M J, Islam N, Ullman HL, Aznar J: Downregulation of human platelet reactivity by neutrophils: Participation of lipoxy- genase derivatives and adhesive proteins. J Clin Invest 92: 1357-1365, 1993

157. Peppelenbosch MR Tertoolen LGJ, Hage W J, de Laat SW: Epidermal growth factor-induced actin remodeling is regu- lated by 5-1ipoxygenase and cyclooxygenase products. Cell 74: 565-575, 1993

158. Kim JA, Gu J, Natarajian R, Berliner JA, Nadler J: Evi- dence that a leukocyte type of 12-1ipoxygenase is expressed in normal human vascular and mononuclear cells. Clin Res 41: 148A, 1993

159. Piomelli D, Volterra A, Dale N, Siegelbaum SA, Kandel ER, Schwartz JH, Belardetti F: Lipoxygenase metabolites of arachidonic acid as second messengers for presynaptic inhibition of Aplysia sensory cells. Nature 328; 38-43,1987

160. Nadler JL, Natarajian R, Stern N: Specifc action of the li- poxygenase pathway in mediating angiotensin II-induced aldosterone synthesis in isolated adrenal glomerula cells. J Clin Invest 80: 1763-1769, 1987

161. Chart CC, Duhamel L, Ford-Hutchison A: Leukotriene B4 and 12-hydroxyeicosatetraenoic acid stimulate epidermal

395

proliferation in vivo in the guinea pig. J Invest Dermatol 85: 333-334, 1985

162. Glasgow WC, Afshari CA, Barret JC, Eling TE: Modula- tion of the epidermal growth factor mitogenic response by metabolites of linoleic and arachidonic acid in Syrian ham- ster embryo fibroblasts. Differential effects in tumor sup- pressor (+) and (-) phenotypes. J Biol Chem 267: 10771- 10779, 1992

163. Antonipillai I: 12-1ipoxygenase products are potent inhib- itors of prostacyclin-induced renin release. Proc Soc Exp Biol Med I94 224-230, 1990

164. Etingin OR, Hajjar DP: Evidence for cytokine regulation of cholesterol metabolism in herpes virus-infected arterial cells by the lipoxygenase pathway. J Lipid Res 31: 299-305, 1990

165. Setty BN, Graeber JE, Stuart M J: The mitogenic effect of 15- and 12-hydroxyeicosatetraenoic acid on endothelial cells may be mediated via diacylglycerol kinase. J Biol Chem 262: 17613-17622, 1987

166. Dethlefsen SM, Shepro D, D'Amore P: Arachidonic acid metabolites in bFGF-, PDGF-, and serum-stimulated vas- cular cell growth. Exp Cell Res 212: 262-273, 1994

167. O'Brian CA, Ward NE: Biology of the protein kinase C family. Cancer Metastasis Rev 8: 19%214, 1989

168. Liu B, Renaud C, Nelson KK, Chen YQ, Bazaz R, Kowynia J, Timar J, Diglio CA, Honn KV: Protein kinase C inhibitor calphostin C reduces B16 amelanotic melanoma cell adhe- sion to endothelium and lung colonization. Int J Cancer 52: 147-152, 1992

169. Gopalakrishina R, Barsky SH: Tumor-promoter-induced membrane-bound protein kinase C regulates hematoge- nous metastasis. Proc Natl Acad Sci USA 85: 612-616,1988

170. Dekker LV, Parker P J: Protein kinase C - a question of specificity. TIBS 19: 73-77, 1994

171. Borner C, Guadagno SN, Hsiao WWL, Fabbro D, Barr M, Weinstein IB: Expression of four protein kinase C isoforms in rat fibroblasts. Differential alterations in ras-, src-, and fos-transformed cells. J Biol Chem 267: 12900-12910, 1992

172. Chun JS, Jacobson BS: Spreading of HeLa cells on a colla- gen substratum requires a second messenger formed by the lipoxygenase metabolism of arachidonic acid released by collagen receptor clustering. Mol Biol Cell 3: 481492,1992

173. Chun JS, Jacobson BS: Requirement for diacylglycerol and protein kinase C in HeLa cell-substratum adhesion and their feedback amplification of arachidonic acid produc- tion for optimum cell spreading. Mol Biol Cell 4: 271-281, 1993

174. Hagerman RA, Smith TJ, Locniskar MF: Lipoxygenase metabolites activate protein kinase C. FASEB J 7: A601, 1993

175. Natarajan R, Lanting L, Nadler XJ: Role of specific iso- forms of protein kinase C in angiotensin II and lipoxyge- nase action in rat adrenal glomerulosa cells. Mol Cell En- docrino1101: 59-66, 1994

176. Berridge M J, Irvine RF: Inositol phosphates and cell sig- naling. Nature 341: 197-205, 1989

396

177. Cho Y, Ziboh VA: 13-hydroxyoctadecaenoic acid reverses epidermal hyperproliferation via selective inhibition of protein kinase C-[~ activity. Biochem Biophys Res Com- mun 201: 257-265, 1994

178. Gardiner PJ: Classification of protanoid receptors. Adv Prostaglandin Thromboxane Leukotriene Res 20: 110-118, 1990

179. Namba T, Oida H, Sugimoto Y, Kakizuka A, Negishi M, Ichikawa A, Narumiya S: cDNA cloning of a mouse prosta- cyclin receptor. Multiple signaling pathways and expres- sion in thymic medulla. J Biol Chem 269: 9986-9992, 1994

180. Bastien L, Sawyer N, Grygorczk R, Metters KM, Adam M: Cloning, functional expression, and characterization of the human prostaglandin E2 receptor EP2 subtype. J Biol Chem 269: 11873-11877, 1994

181. Boie Y, Rushmore T, Darmon-Goodwin A, Grygorczk R, Slipetz DM, Metters KM, Abramovitz M: Cloning and ex- pression of a cDNA for the human prostanoid IP receptor. J Biol Chem 269: 12173-12178, 1994

182. Gross E, Ruzicka T, Restorff BV, Stolz W, Klotz K-N: High-affinity binding and lack of growth-promoting activ- ity of 12(S)-hydroxyeicosatetraenoic acid (12(S)-HETE) in a human epidermal cell line. J Invest Dermato194: 446-451, 1990

183. Croset M, Lagarde M: Stereospecific inhibition of PGH2- induced aggregation by lipoxygenase products of eicosae- noic acids. Biochem Biophys Res Commun 112: 878-883, 1983

184. Fonlupt R Croset M, Lagarde M: 12(S)-HETE inhibits the binding of PGH2/TXA 2 receptor ligands in human plate- lets. Thromb Res 63: 239-248, 1991

185. Herbertsson H, Hammarstrom S: High-affinity binding sites for 12(S)-hydroxy-5,8,10,14-eicosatetraenoic acid (12(S)-HETE) in carcinoma cells. FEB8 Lett 298: 249-252, 1992

186. Herbertsson H, Hammarstrom S: Cytosolic 12(8)-hy- droxy-5,8,10,14-eicosatetraenoic acid binding sites in carci- noma cells. In: Honn KV, Marnett LJ, Nigam S (eds) Eico-

sanoids and Other Bioactive Lipids in Cancer, Inflamma- tion and Radiation Injury. Kluwer, Boston, 1994, in press

187. Kohn EC, Liotta LA: Invasion and metastasis: An old problem with new approaches. Oncology 7: 47-52, 1993

188. Kohn EC, Sandeen MA, Liotta LA: In vivo efficacy of a novel inhibitor of selected signal transduction pathways in- cluding calcium, arachidonate, and inositol phosphates. Cancer Res 52: 3208-3212, 1992

189. Hagmann W, Gao X, Timar J, Chen YQ, Fahrenkopf C, Kagawa D, Lee M, Zacharek A, Honn KV: Intracellular localization, activity, and bidirectional modulation of 12- lipoxygenase in A431 epidermoid carcinoma cells: Influen- ce of epidermal growth factor, starvation, and inhibition of protein tyrosine kinase. Submitted

190. Silletti S, Timar J, Honn KV, Raz A: Autocrine motility fac- tor induces 12-1ipoxygenase expression and activity in high- but not low-metastatic murine melanoma cells. Can- cer Res, 1994, in press

191. Eggens I, Fenderson BA, Toyokuni T, Dean B, Stroud MR, Hakomori S: Specific interaction between Le x and Le × de- terminants: a possible basis for cell recognition in preim- plantation embryos and in embryonal carcinoma cells. J Biol Chem 264: 9476-9484, 1989

192. Hakomori S: New directions in cancer therapy based on abrrant expression of glycosphingolipids: anti-adhesion and ortho-signaling therapy. Cancer Cells 3: 461470, 1991

193, Liu B, Khan WA, Hannun YA, Bazaz R, Renaud C, Stoja- kovic S, Timar J, Taylor J, Honn KV: 12(S)-HETE and 13(S)-HODE regulation of protein kinase C alpha in mela- noma cells: Role of receptor mediated hydrolysis of inosi- tol phospholipids. Submitted

194. Liu B, Honn KV: unpublished observations 195. Tang DG, Honn KV: 12-Lipoxygenase, 12(S)-HETE, and

Cancer Metastasis. Annals New York Acad Sci 1994, in press

Address for offprints: K.V. Honn, Division of Cancer Biology, Department of Radiation Oncology, Wayne State University, 431 Chemistry, Detroit, M148202, USA