Abstract

Galectins are a family of carbohydrate-binding proteins with an affinity for β-galactosides. Galectin-1 (Gal-1) is differentially expressed by various normal and pathological tissues and appears to be functionally polyvalent, with a wide range of biological activity. The intracellular and extracellular activity of Gal-1 has been described. Evidence points to Gal-1 and its ligands as one of the master regulators of such immune responses as T-cell homeostasis and survival, T-cell immune disorders, inflammation and allergies as well as host–pathogen interactions. Gal-1 expression or overexpression in tumors and/or the tissue surrounding them must be considered as a sign of the malignant tumor progression that is often related to the long-range dissemination of tumoral cells (metastasis), to their dissemination into the surrounding normal tissue, and to tumor immune-escape. Gal-1 in its oxidized form plays a number of important roles in the regeneration of the central nervous system after injury. The targeted overexpression (or delivery) of Gal-1 should be considered as a method of choice for the treatment of some kinds of inflammation-related diseases, neurodegenerative pathologies and muscular dystrophies. In contrast, the targeted inhibition of Gal-1 expression is what should be developed for therapeutic applications against cancer progression. Gal-1 is thus a promising molecular target for the development of new and original therapeutic tools.

Galectins: an overview

Galectins are a phylogenetically conserved family of lectins defined in 1994 as a shared consensus of amino-acid-sequences of about 130 amino acids and the carbohydrate recognition domain (CRD) responsible for β-galactoside binding (Barondes et al., 1994). Fifteen mammalian galectins have been identified to date. While some of these galectins contain one CRD and are biologically active as monomers (galectins-5, -7, -10), as homodimers (galectins-1, -2, -11, 13–14, -15) or as oligomers that aggregate though their non-lectin domain (galectin-3); others contain two CRDs connected by a short linker peptide (galectins-4, -6, -8, -9, -12). While the CRDs of all the galectins share an affinity for the minimum saccharide ligand N-acetyllactosamine—a common disaccharide found on many cellular glycoproteins—individual galectins can also recognize different modifications to this minimum saccharide ligand and so demonstrate the fine specificity of certain galectins for tissue- or developmentally-specific ligands (Ahmad et al., 2004). Location studies of galectins have established that these proteins can segregate into multiple cell compartments in function of the status of the cells in question (Danguy et al., 2002; Liu and Rabinovich, 2005). Although galectins as a whole do not have the signal sequence required for protein secretion through the usual secretory pathway, some galectins are secreted and are found in the extracellular space (Hughes, 1999). While the intracellular activity of galectin-1 (Gal-1) is mainly independent on its lectin activity, its extracellular activity is mainly dependent on it.

Gal-1: molecular structures at gene and protein levels

The first protein discovered in the family was Gal-1. As reported by the MapViewer program and the Entrez genome database on the NCBI website (http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?db=Genome&itool=toolbar), Gal-1 is encoded by the LSGALS1 gene located on chromosome 22q12 (Figure 1). The 0.6 kb transcript (GenBank: NM_002305) is the result of the splicing of four exons encoding a protein with 135 amino acids (GenPept : NP_002296, SwissProt: P09382) (Figure 2). Gal-1 occurs as a monomer as well as a non-covalent homodimer consisting of subunits of one CRD (Gal-1, ∼29 kDa) (Barondes et al., 1994; Cho and Cummings, 1995). Each form is associated with different biological activities, as detailed below.

Fig. 1.

Map of the Gal-1 gene on the human chromosome 22q12. The sequences were retrieved using the MapViewer program and the Entrez genome database on the NCBI website and were analyzed using the GeneWorks program produced by IntelliGenetics Inc. The curved arrows indicate the initial transcription start-up sites (Chiariotti et al., 2004). The 0.6 kb transcript results from the splicing of four exons (the boxes indicate the four coding regions) and encodes for a protein of 135 aminoacids.

Fig. 2.

Structure of Gal-1 protein. The overall folding of Gal-1 involves a β sandwich consisting of two anti-parallel β-sheets of five (F1–F5) and six (S1–S6a/b) strands respectively. The N and the C termini of each monomer are positioned at the dimer interface and the CRDs are located at the far ends of the same face, a configuration which constitutes a long, negatively charged cleft in the cavity (Lopez-Lucendo et al., 2004). The green amino acid symbols illustrate highly conserved residues. The key residues of the CRD, which is known to interact directly with bound carbohydrate by means of hydrogen bounds, are colored pink, while those interacting with carbohydrates via van der Waals interactions are orange (Cooper, 2002) and include His44, Asn46, Arg48, Val59, Asn61, Trp68, Glu71, and Arg73. The 3D ribbon diagram of the homodimeric human Gal-1 was designed with MOLSCRIPT by Lopez-Lucendo et al. (2004). A lactose (Galβ1–4Glc) is illustrated in the CRD.

The overall folding of Gal-1 involves a β sandwich consisting of two anti-parallel β-sheets (Figure 2). This jellyroll topology of the CRD constitutes the typical folding patterns of galectins. Human Gal-1 exists as a dimer in solution (Lopez-Lucendo et al., 2004). The integrity of this dimer is maintained principally by interactions between the monomers at the interface and through the well-conserved hydrophobic core, a factor which explains the observed stability of the dimer in molecular terms (Figure 2) (Lopez-Lucendo et al., 2004). Nevertheless, one of the main characteristics of the homodimeric Gal-1 protein is that it spontaneously dissociates at low concentrations (Kd ∼ 7 µM) into a monomeric form that is still able to bind to carbohydrates (Cho and Cummings, 1995), but with a lower level of affinity (Leppanen et al., 2005). In the present paper Gal-1 will be used as the general term for both the monomeric and the dimeric product of the LGALS1 gene. We will clearly state monomeric (mGal-1) or dimeric galectin-1 (dGal-1) when reporting results specific to the monomer or the dimer form respectively. Gal-1 can also exist in an oxidized form, that is, a form that lacks lectin activity (Outenreath and Jones, 1992).

The regulation of Gal-1 expression

A small region spanning the initial transcription start-up site (–63/+45) is sufficient in the promoter region of the Gal-1 gene for its transcriptional activity in mice (Chiariotti et al., 2004) (Figure 1). Both an upstream and a downstream position-dependent cis-element are necessary for efficient transcriptional activity; an additional start-up site has been mapped at position-31, and an Sp1-binding site (–57/–48) and a consensus initiator (Inr) element (which partially overlaps a non-canonical TATA box) direct RNA initiation (Chiariotti et al., 2004) (Figure 1). The upstream transcripts contribute to more than half of the Gal-1 mRNA population (Figure 1). The 5′- end of this transcript is extremely GC-rich and may fold into a stable hairpin structure which could influence translation (Chiariotti et al., 2004). The approximate position of the other putative and/or characterized regulatory elements is indicated in Figure 1 and relates to the CAAT box, to the nuclear factor kappaB-binding site (NF-κB) and to the sodium butyrate and retinoic acid (RA) response sequences. The various physicochemical agents already known as being able to modulate the expression of Gal-1 are listed in Table I. The methylation status of the Gal-1 promoter is also a very important mechanism that controls the expression of the gene (Chiariotti et al., 2004).

Table I.

The regulation of Gal-1 expression

EffectorsTypesDosesTissuesEffectsReferences
5-AzacytidineDemethylation of promoterHepatocellular carcinomas (hu), liver, thyroid cells (rat), osteosarcoma cells (hu)Kondoh et al., 2003; Chiariotti et al., 2004
Irreversible promoter demethylation2 µMT leukemia cells (hu)Poirier et al., 2001; Chiariotti et al., 2004
10 µMB lymphomas (hu)
BenzodiazepinePeripheral benzodiazepine receptorJurkat T lymphoma cells (hu)Rochard et al., 2004
Budesonide glucocorticoidAnti-inflammatory250 ng/mLNasal polyps (hu)Delbrouck et al., 2002
CyclophosphamideAntimetastatic alkylating drugLow dose, 10 mg/kgLymphomas (rat)Rabinovich, Rubinstein et al., 2002
ERBB2 overexpressionHER2/neu oncogeneBreast cancer cells (hu)Mackay et al., 2003
Hepatitis C virus core protein/ethanolMAPK (ERK/p38) activationLiver (mo)Tsutsumi et al., 2003
Imp (−/−) miceInsulin-like growth factor II mrRNA- binding protein 1 (IMP1)Impaired gut development (mo)Hansen et al., 2004
Minimally oxidized low-density lipoprotein (MM-LDL)Pro(?)-inflammatory circulating lipoprotein100 µg/mLEndothelial cells (hu)Baum et al., 1995; Perillo et al., 1995
ProgesteroneHormonesReceptor antagonist blocks gal-1 expressionUterine tissue (mo)Choe et al., 1997
EstrogenUterine tissue (mo)Choe et al., 1997
Estrogen/delta fosBRat1a embryo cells (rat)Tahara et al., 2003
Thyroid stimulating hormone (TSH)1 nMThyroid (rat)Chiariotti et al., 1994
Retinoic acid (all-trans)Differentiating1 µMEmbryonal carcinomas, myoblastic cells (mo)Lu et al., 2000
Transformed neural cells (rat)Chiariotti et al., 1994
Cholesteatomas (hu)Relation with level of receptor expressionSimon et al., 2001
Sodium butyrateDifferentiating1–4 mMColon carcinomas (hu)Ohannesian et al., 1994
3 mMHNSCCs (hu)Gillenwater et al., 1998
Embryonal carcinomas (mo)Lu and Lotan, 1999
2–5 mMGlobelet cells (hu)Gaudier et al., 2004
2.5–10 mMProstate cancers (hu)Ellerhorst, Nguyen, Cooper, Estrov et al., 1999
Valproic acidAntiepileptic drug, inducer of neural tube defectsEmbryo (mo)Kultima et al., 2004
EffectorsTypesDosesTissuesEffectsReferences
5-AzacytidineDemethylation of promoterHepatocellular carcinomas (hu), liver, thyroid cells (rat), osteosarcoma cells (hu)Kondoh et al., 2003; Chiariotti et al., 2004
Irreversible promoter demethylation2 µMT leukemia cells (hu)Poirier et al., 2001; Chiariotti et al., 2004
10 µMB lymphomas (hu)
BenzodiazepinePeripheral benzodiazepine receptorJurkat T lymphoma cells (hu)Rochard et al., 2004
Budesonide glucocorticoidAnti-inflammatory250 ng/mLNasal polyps (hu)Delbrouck et al., 2002
CyclophosphamideAntimetastatic alkylating drugLow dose, 10 mg/kgLymphomas (rat)Rabinovich, Rubinstein et al., 2002
ERBB2 overexpressionHER2/neu oncogeneBreast cancer cells (hu)Mackay et al., 2003
Hepatitis C virus core protein/ethanolMAPK (ERK/p38) activationLiver (mo)Tsutsumi et al., 2003
Imp (−/−) miceInsulin-like growth factor II mrRNA- binding protein 1 (IMP1)Impaired gut development (mo)Hansen et al., 2004
Minimally oxidized low-density lipoprotein (MM-LDL)Pro(?)-inflammatory circulating lipoprotein100 µg/mLEndothelial cells (hu)Baum et al., 1995; Perillo et al., 1995
ProgesteroneHormonesReceptor antagonist blocks gal-1 expressionUterine tissue (mo)Choe et al., 1997
EstrogenUterine tissue (mo)Choe et al., 1997
Estrogen/delta fosBRat1a embryo cells (rat)Tahara et al., 2003
Thyroid stimulating hormone (TSH)1 nMThyroid (rat)Chiariotti et al., 1994
Retinoic acid (all-trans)Differentiating1 µMEmbryonal carcinomas, myoblastic cells (mo)Lu et al., 2000
Transformed neural cells (rat)Chiariotti et al., 1994
Cholesteatomas (hu)Relation with level of receptor expressionSimon et al., 2001
Sodium butyrateDifferentiating1–4 mMColon carcinomas (hu)Ohannesian et al., 1994
3 mMHNSCCs (hu)Gillenwater et al., 1998
Embryonal carcinomas (mo)Lu and Lotan, 1999
2–5 mMGlobelet cells (hu)Gaudier et al., 2004
2.5–10 mMProstate cancers (hu)Ellerhorst, Nguyen, Cooper, Estrov et al., 1999
Valproic acidAntiepileptic drug, inducer of neural tube defectsEmbryo (mo)Kultima et al., 2004

Hu, human; mo, mouse; ↓, decreases and/or inhibits; ↑, increases and/or favors; ?, uncertain.

Table I.

The regulation of Gal-1 expression

EffectorsTypesDosesTissuesEffectsReferences
5-AzacytidineDemethylation of promoterHepatocellular carcinomas (hu), liver, thyroid cells (rat), osteosarcoma cells (hu)Kondoh et al., 2003; Chiariotti et al., 2004
Irreversible promoter demethylation2 µMT leukemia cells (hu)Poirier et al., 2001; Chiariotti et al., 2004
10 µMB lymphomas (hu)
BenzodiazepinePeripheral benzodiazepine receptorJurkat T lymphoma cells (hu)Rochard et al., 2004
Budesonide glucocorticoidAnti-inflammatory250 ng/mLNasal polyps (hu)Delbrouck et al., 2002
CyclophosphamideAntimetastatic alkylating drugLow dose, 10 mg/kgLymphomas (rat)Rabinovich, Rubinstein et al., 2002
ERBB2 overexpressionHER2/neu oncogeneBreast cancer cells (hu)Mackay et al., 2003
Hepatitis C virus core protein/ethanolMAPK (ERK/p38) activationLiver (mo)Tsutsumi et al., 2003
Imp (−/−) miceInsulin-like growth factor II mrRNA- binding protein 1 (IMP1)Impaired gut development (mo)Hansen et al., 2004
Minimally oxidized low-density lipoprotein (MM-LDL)Pro(?)-inflammatory circulating lipoprotein100 µg/mLEndothelial cells (hu)Baum et al., 1995; Perillo et al., 1995
ProgesteroneHormonesReceptor antagonist blocks gal-1 expressionUterine tissue (mo)Choe et al., 1997
EstrogenUterine tissue (mo)Choe et al., 1997
Estrogen/delta fosBRat1a embryo cells (rat)Tahara et al., 2003
Thyroid stimulating hormone (TSH)1 nMThyroid (rat)Chiariotti et al., 1994
Retinoic acid (all-trans)Differentiating1 µMEmbryonal carcinomas, myoblastic cells (mo)Lu et al., 2000
Transformed neural cells (rat)Chiariotti et al., 1994
Cholesteatomas (hu)Relation with level of receptor expressionSimon et al., 2001
Sodium butyrateDifferentiating1–4 mMColon carcinomas (hu)Ohannesian et al., 1994
3 mMHNSCCs (hu)Gillenwater et al., 1998
Embryonal carcinomas (mo)Lu and Lotan, 1999
2–5 mMGlobelet cells (hu)Gaudier et al., 2004
2.5–10 mMProstate cancers (hu)Ellerhorst, Nguyen, Cooper, Estrov et al., 1999
Valproic acidAntiepileptic drug, inducer of neural tube defectsEmbryo (mo)Kultima et al., 2004
EffectorsTypesDosesTissuesEffectsReferences
5-AzacytidineDemethylation of promoterHepatocellular carcinomas (hu), liver, thyroid cells (rat), osteosarcoma cells (hu)Kondoh et al., 2003; Chiariotti et al., 2004
Irreversible promoter demethylation2 µMT leukemia cells (hu)Poirier et al., 2001; Chiariotti et al., 2004
10 µMB lymphomas (hu)
BenzodiazepinePeripheral benzodiazepine receptorJurkat T lymphoma cells (hu)Rochard et al., 2004
Budesonide glucocorticoidAnti-inflammatory250 ng/mLNasal polyps (hu)Delbrouck et al., 2002
CyclophosphamideAntimetastatic alkylating drugLow dose, 10 mg/kgLymphomas (rat)Rabinovich, Rubinstein et al., 2002
ERBB2 overexpressionHER2/neu oncogeneBreast cancer cells (hu)Mackay et al., 2003
Hepatitis C virus core protein/ethanolMAPK (ERK/p38) activationLiver (mo)Tsutsumi et al., 2003
Imp (−/−) miceInsulin-like growth factor II mrRNA- binding protein 1 (IMP1)Impaired gut development (mo)Hansen et al., 2004
Minimally oxidized low-density lipoprotein (MM-LDL)Pro(?)-inflammatory circulating lipoprotein100 µg/mLEndothelial cells (hu)Baum et al., 1995; Perillo et al., 1995
ProgesteroneHormonesReceptor antagonist blocks gal-1 expressionUterine tissue (mo)Choe et al., 1997
EstrogenUterine tissue (mo)Choe et al., 1997
Estrogen/delta fosBRat1a embryo cells (rat)Tahara et al., 2003
Thyroid stimulating hormone (TSH)1 nMThyroid (rat)Chiariotti et al., 1994
Retinoic acid (all-trans)Differentiating1 µMEmbryonal carcinomas, myoblastic cells (mo)Lu et al., 2000
Transformed neural cells (rat)Chiariotti et al., 1994
Cholesteatomas (hu)Relation with level of receptor expressionSimon et al., 2001
Sodium butyrateDifferentiating1–4 mMColon carcinomas (hu)Ohannesian et al., 1994
3 mMHNSCCs (hu)Gillenwater et al., 1998
Embryonal carcinomas (mo)Lu and Lotan, 1999
2–5 mMGlobelet cells (hu)Gaudier et al., 2004
2.5–10 mMProstate cancers (hu)Ellerhorst, Nguyen, Cooper, Estrov et al., 1999
Valproic acidAntiepileptic drug, inducer of neural tube defectsEmbryo (mo)Kultima et al., 2004

Hu, human; mo, mouse; ↓, decreases and/or inhibits; ↑, increases and/or favors; ?, uncertain.

The subcellular distribution and the export of Gal-1

Gal-1 is present both inside and outside cells and has both intracellular and extracellular functions (Figure 3). Gal-1 shows the characteristics of typical cytoplasmic proteins as well as an acetylated N-terminus and a lack of glycosylations (Clerch et al., 1988) (Figure 2); it has been described in cell nuclei and cytosols and also translocates to the intracellular side of cell membranes (Figure 3). Nevertheless, even though Gal-1 lacks recognizable secretion signal sequences and does not pass along the standard endoplasmic reticulum/Golgi pathway (Hughes, 1999), it is well-known that it is secreted and can be found on the extracellular side of all cell membranes as well as in the extracellular matrices of various normal and neoplastic tissues (Cooper and Barondes, 1990; van den Brule et al., 1997, 2003; Clausse et al., 1999; Camby et al., 2002; Danguy et al., 2002; von Wolff et al., 2005). There is evidence that this protein is secreted in a manner similar to fibroblast growth factor-2 (FGF-2) (Nickel, 2005) via inside-out transportation involving direct translocation across the plasma membrane of mammalian cells and requiring as yet unidentified integral membrane proteins and cytosolic factors (Nickel, 2005). The β-galactoside-binding site may constitute the primary targeting motif for galectin export machinery using β-galactoside-containing surface molecules as export receptors for intracellular Gal-1 (Nickel, 2005). As a consequence there is a quality control mechanism present since the export machinery recognizes only properly folded Gal-1. The similarity of the FGF-2 and Gal-1 export pathways suggests an important role for the sodium pump (the Na+/K+–ATPase) in their export features because ouabain, a selective inhibitor of the sodium pump, inhibits these export processes (Nickel, 2005).

Fig. 3.

Gal-1 and cell signaling. Gal-1 is present both inside and outside cells, and has both intracellular and extracellular functions. The extracellular functions require the carbohydrate-binding properties of dGal-1 while the intracellular ones are associated with carbohydrate-independent interactions between Gal-1 and other proteins. The Gal-1 induced growth inhibition requires functional interactions with the α5β1 integrin (Fischer et al., 2005). The antiproliferative effects result from the inhibition of the Ras-MEK-ERK pathway and the consecutive transcriptional induction of p27: two Sp1-binding sites in the p27 promoter are crucial to Gal-1 responsiveness (Fischer et al., 2005). The inhibition of the Ras-MEK-ERK cascade by Gal-1 increases Sp1 transactivation, with DNA binding relating to the reduced threonine-phosphorylation of Sp1 (Fischer et al., 2005). Furthermore, Gal-1 induces p21 transcription and selectively increases p27 protein stability. The Gal-1-mediated accumulation of p27 and p21 inhibits Cdk2 activity and, ultimately, results in G1 cell cycle arrest and growth inhibition (Fischer et al., 2005). The Gal-1-induced increase in cell motility involves the Gal-1-induced increase in rhoA expression and the alteration of the polymerization of the actin cytoskeleton (Camby et al., 2002). Gal-1 is recruited from the cytosol to the cell membrane by H-Ras-GTP in a lactose-independent manner with the resulting stabilization of the H-Ras-GTP, the clustering of the H-RAS-GTP and Gal-1 in non-raft microdomains (Prior et al., 2003), the subsequent binding to Raf-1 (but not to PI3Kinase) and the activation of the ERK signaling pathway (Elad-Sfadia et al., 2002). Nuclear Gal-1 interacts with Gemin4 and is co-immunoprecipitated with the nuclear SMN complexes involved in the splicing pathway (Vyakarnam et al., 1997).

Gal-1 binding partners

Although galectins in general, and Gal-1 in particular, were first described as lectins that bind β-galactosides, it is now clear from the literature that as well as being a lectin, Gal-1 is also engaged in protein–protein interactions (Table II). Interestingly enough, in most cases the lectin activity of Gal-1 is observed when it is extracellular, while the protein–protein interactions of Gal-1 concern its intracellular functions. The fact nevertheless remains that one can wonder what is the real biological relevance of such a high number of Gal-1 binding partners (Table II). Binding affinity studies are warranted in order to determine the potential binding partners which are real for Gal-1 and the ones which are not because they are associated with excessively weak affinities.

Table II.

The binding partners of Gal-1

Binding partnersMonomeric/dimeric Gal-1Binding type (P–C, P–P)Cell/tissue typesBiological functionsReferences
β1 integrinDimeric
    α1β1, α7β1P–CSkeletal and vascular SMCAdhesion, FAK activationGu et al., 1994; Moiseeva et al., 1999; Moiseeva, Williams, Goodall et al., 2003
    α5β1P–CEpithelial carcinoma cellsInhibit ras-MEK-ERK pathway, increase p21 and p27, and growth inhibitionFischer et al., 2005
    αMβ2 integrinP–CMacrophage, neutrophils (?)NS activationAvni et al., 1998; Almkvist et al., 2002
1B2 glycolipidNSP–COlfactory axon in olfactory bulb↑ cell–cell and cell–laminin adhesionMahanthappa et al., 1994
ActinNSP–PBrainNSJoubert et al., 1992
P–C (?)MOLT-4 T cellsPace et al., 1999
CA-125NSP–CHeLa cellsGal-1 export (?)Seelenmeyer et al., 2003
CD2/CD3DimericP–CJurkat T cellsMembrane redistribution, induction of cell deathPace et al., 1999; Walzel et al., 2000
CD4DimericP–CT cellNSPace et al., 1999; 2000
CD43DimericP–CT cellsMembrane redistribution, induction of cell deathPace et al., 1999; Nguyen et al., 2001; Lanteri et al., 2003
CD45DimericP–CT, B cellsMembrane redistribution, induction of cell deathPerillo et al., 1995; Fouillit et al., 2000; Symons et al., 2000; Fajka-Boja et al., 2002
CD7NSP–CT cellsInduction of cell deathPace et al., 1999, 2000
Carcino embryonic antigen (CEA, CD66e)NSP–CKM12 colon carcinoma cellsNSOhannesian et al., 1994
Cytochrome oxidase subunit IIINSP–P (?)HeLa cellsNSPark et al., 2001
FibronectinNSP–COvarian carcinoma, placenta↑ adhesionOzeki et al., 1995; Moiseeva et al., 2000; Moiseeva, Williams, and Samani, 2003; van den Brule et al., 2003
Genim-4 nuclear and (?) cytoplasmicNSP–PHeLa cellspreRNA splicing, RNA interferencePark et al., 2001; Hutvagner and Zamore, 2002
Glycoprotein 90K (MAC-2BP)NSP–CA375 melanoma cells↑ cell aggregationTinari et al., 2001
Glycosaminoglycan (chondroitin sulphate B, heparan sulfate)NSP–CVSMCModulation of ECM assembly, ↓ adhesionMoiseeva, Williams, Goodall et al., 2003
GM1 gangliosideDimericP–CSK-N-MC neurobastoma cells↓ growthKopitz et al., 1998, 2001; Andre et al., 2004
HBGp82NSP–CBrainNSChadli et al., 1997
H-rasDimericP–PHeLa, HEK293, Rat‐1, 293T cells↑ ras activation with selective activation of Raf-1/ERK pathwayPaz et al., 2001; Elad-Sfadia et al., 2002; Prior et al., 2003; Rotblat et al., 2004
LamininNSP–CMelanomas, myoblasts, ovarian carcinomas, Leydig cells, placenta↑ adhesionOzeki et al., 1995; Moiseeva et al., 2000; Moiseeva, Williams, Goodall et al., 2003; van den Brule et al., 2003; Martinez et al., 2004
LAMP-1 (CD107a), LAMP-2 (CD107b)NSP–COvarian, colon carcinomas↑ adhesionOhannesian et al., 1994; Woynarowska et al., 1994
MucinNSP–CEpithelial glycocalyces of gastric and intestinal mucosaNSWasano and Hirakawa, 1997
OsteospontinNSP–CVSMC↑ adhesionMoiseeva et al., 2000
Pre-B cell receptorNSP–CB cell lines↑ adhesion, cell diffrentiationGauthier et al., 2002
SUMO-3/SMT3BNSP–P (?)HeLa cellsNSPark et al., 2001
ThrombospondinNSP–CVSMC↑ adhesionMoiseeva et al., 2000
Thy-1NSP–CT cellsNSSymons et al., 2000
VitronectinNSP–CVSMCECM assemblyMoiseeva, Williams, and Samani, 2003
Binding partnersMonomeric/dimeric Gal-1Binding type (P–C, P–P)Cell/tissue typesBiological functionsReferences
β1 integrinDimeric
    α1β1, α7β1P–CSkeletal and vascular SMCAdhesion, FAK activationGu et al., 1994; Moiseeva et al., 1999; Moiseeva, Williams, Goodall et al., 2003
    α5β1P–CEpithelial carcinoma cellsInhibit ras-MEK-ERK pathway, increase p21 and p27, and growth inhibitionFischer et al., 2005
    αMβ2 integrinP–CMacrophage, neutrophils (?)NS activationAvni et al., 1998; Almkvist et al., 2002
1B2 glycolipidNSP–COlfactory axon in olfactory bulb↑ cell–cell and cell–laminin adhesionMahanthappa et al., 1994
ActinNSP–PBrainNSJoubert et al., 1992
P–C (?)MOLT-4 T cellsPace et al., 1999
CA-125NSP–CHeLa cellsGal-1 export (?)Seelenmeyer et al., 2003
CD2/CD3DimericP–CJurkat T cellsMembrane redistribution, induction of cell deathPace et al., 1999; Walzel et al., 2000
CD4DimericP–CT cellNSPace et al., 1999; 2000
CD43DimericP–CT cellsMembrane redistribution, induction of cell deathPace et al., 1999; Nguyen et al., 2001; Lanteri et al., 2003
CD45DimericP–CT, B cellsMembrane redistribution, induction of cell deathPerillo et al., 1995; Fouillit et al., 2000; Symons et al., 2000; Fajka-Boja et al., 2002
CD7NSP–CT cellsInduction of cell deathPace et al., 1999, 2000
Carcino embryonic antigen (CEA, CD66e)NSP–CKM12 colon carcinoma cellsNSOhannesian et al., 1994
Cytochrome oxidase subunit IIINSP–P (?)HeLa cellsNSPark et al., 2001
FibronectinNSP–COvarian carcinoma, placenta↑ adhesionOzeki et al., 1995; Moiseeva et al., 2000; Moiseeva, Williams, and Samani, 2003; van den Brule et al., 2003
Genim-4 nuclear and (?) cytoplasmicNSP–PHeLa cellspreRNA splicing, RNA interferencePark et al., 2001; Hutvagner and Zamore, 2002
Glycoprotein 90K (MAC-2BP)NSP–CA375 melanoma cells↑ cell aggregationTinari et al., 2001
Glycosaminoglycan (chondroitin sulphate B, heparan sulfate)NSP–CVSMCModulation of ECM assembly, ↓ adhesionMoiseeva, Williams, Goodall et al., 2003
GM1 gangliosideDimericP–CSK-N-MC neurobastoma cells↓ growthKopitz et al., 1998, 2001; Andre et al., 2004
HBGp82NSP–CBrainNSChadli et al., 1997
H-rasDimericP–PHeLa, HEK293, Rat‐1, 293T cells↑ ras activation with selective activation of Raf-1/ERK pathwayPaz et al., 2001; Elad-Sfadia et al., 2002; Prior et al., 2003; Rotblat et al., 2004
LamininNSP–CMelanomas, myoblasts, ovarian carcinomas, Leydig cells, placenta↑ adhesionOzeki et al., 1995; Moiseeva et al., 2000; Moiseeva, Williams, Goodall et al., 2003; van den Brule et al., 2003; Martinez et al., 2004
LAMP-1 (CD107a), LAMP-2 (CD107b)NSP–COvarian, colon carcinomas↑ adhesionOhannesian et al., 1994; Woynarowska et al., 1994
MucinNSP–CEpithelial glycocalyces of gastric and intestinal mucosaNSWasano and Hirakawa, 1997
OsteospontinNSP–CVSMC↑ adhesionMoiseeva et al., 2000
Pre-B cell receptorNSP–CB cell lines↑ adhesion, cell diffrentiationGauthier et al., 2002
SUMO-3/SMT3BNSP–P (?)HeLa cellsNSPark et al., 2001
ThrombospondinNSP–CVSMC↑ adhesionMoiseeva et al., 2000
Thy-1NSP–CT cellsNSSymons et al., 2000
VitronectinNSP–CVSMCECM assemblyMoiseeva, Williams, and Samani, 2003

FAK, facal adhesion kinase; P–C, protein–carbohydrate interaction; P–P, protein–protein interaction; NS, not specified; VSMC, vascular smooth muscle cell; ↓, decreases and/or inhibits; ↑, increases and/or favors; ?, uncertain.

Table II.

The binding partners of Gal-1

Binding partnersMonomeric/dimeric Gal-1Binding type (P–C, P–P)Cell/tissue typesBiological functionsReferences
β1 integrinDimeric
    α1β1, α7β1P–CSkeletal and vascular SMCAdhesion, FAK activationGu et al., 1994; Moiseeva et al., 1999; Moiseeva, Williams, Goodall et al., 2003
    α5β1P–CEpithelial carcinoma cellsInhibit ras-MEK-ERK pathway, increase p21 and p27, and growth inhibitionFischer et al., 2005
    αMβ2 integrinP–CMacrophage, neutrophils (?)NS activationAvni et al., 1998; Almkvist et al., 2002
1B2 glycolipidNSP–COlfactory axon in olfactory bulb↑ cell–cell and cell–laminin adhesionMahanthappa et al., 1994
ActinNSP–PBrainNSJoubert et al., 1992
P–C (?)MOLT-4 T cellsPace et al., 1999
CA-125NSP–CHeLa cellsGal-1 export (?)Seelenmeyer et al., 2003
CD2/CD3DimericP–CJurkat T cellsMembrane redistribution, induction of cell deathPace et al., 1999; Walzel et al., 2000
CD4DimericP–CT cellNSPace et al., 1999; 2000
CD43DimericP–CT cellsMembrane redistribution, induction of cell deathPace et al., 1999; Nguyen et al., 2001; Lanteri et al., 2003
CD45DimericP–CT, B cellsMembrane redistribution, induction of cell deathPerillo et al., 1995; Fouillit et al., 2000; Symons et al., 2000; Fajka-Boja et al., 2002
CD7NSP–CT cellsInduction of cell deathPace et al., 1999, 2000
Carcino embryonic antigen (CEA, CD66e)NSP–CKM12 colon carcinoma cellsNSOhannesian et al., 1994
Cytochrome oxidase subunit IIINSP–P (?)HeLa cellsNSPark et al., 2001
FibronectinNSP–COvarian carcinoma, placenta↑ adhesionOzeki et al., 1995; Moiseeva et al., 2000; Moiseeva, Williams, and Samani, 2003; van den Brule et al., 2003
Genim-4 nuclear and (?) cytoplasmicNSP–PHeLa cellspreRNA splicing, RNA interferencePark et al., 2001; Hutvagner and Zamore, 2002
Glycoprotein 90K (MAC-2BP)NSP–CA375 melanoma cells↑ cell aggregationTinari et al., 2001
Glycosaminoglycan (chondroitin sulphate B, heparan sulfate)NSP–CVSMCModulation of ECM assembly, ↓ adhesionMoiseeva, Williams, Goodall et al., 2003
GM1 gangliosideDimericP–CSK-N-MC neurobastoma cells↓ growthKopitz et al., 1998, 2001; Andre et al., 2004
HBGp82NSP–CBrainNSChadli et al., 1997
H-rasDimericP–PHeLa, HEK293, Rat‐1, 293T cells↑ ras activation with selective activation of Raf-1/ERK pathwayPaz et al., 2001; Elad-Sfadia et al., 2002; Prior et al., 2003; Rotblat et al., 2004
LamininNSP–CMelanomas, myoblasts, ovarian carcinomas, Leydig cells, placenta↑ adhesionOzeki et al., 1995; Moiseeva et al., 2000; Moiseeva, Williams, Goodall et al., 2003; van den Brule et al., 2003; Martinez et al., 2004
LAMP-1 (CD107a), LAMP-2 (CD107b)NSP–COvarian, colon carcinomas↑ adhesionOhannesian et al., 1994; Woynarowska et al., 1994
MucinNSP–CEpithelial glycocalyces of gastric and intestinal mucosaNSWasano and Hirakawa, 1997
OsteospontinNSP–CVSMC↑ adhesionMoiseeva et al., 2000
Pre-B cell receptorNSP–CB cell lines↑ adhesion, cell diffrentiationGauthier et al., 2002
SUMO-3/SMT3BNSP–P (?)HeLa cellsNSPark et al., 2001
ThrombospondinNSP–CVSMC↑ adhesionMoiseeva et al., 2000
Thy-1NSP–CT cellsNSSymons et al., 2000
VitronectinNSP–CVSMCECM assemblyMoiseeva, Williams, and Samani, 2003
Binding partnersMonomeric/dimeric Gal-1Binding type (P–C, P–P)Cell/tissue typesBiological functionsReferences
β1 integrinDimeric
    α1β1, α7β1P–CSkeletal and vascular SMCAdhesion, FAK activationGu et al., 1994; Moiseeva et al., 1999; Moiseeva, Williams, Goodall et al., 2003
    α5β1P–CEpithelial carcinoma cellsInhibit ras-MEK-ERK pathway, increase p21 and p27, and growth inhibitionFischer et al., 2005
    αMβ2 integrinP–CMacrophage, neutrophils (?)NS activationAvni et al., 1998; Almkvist et al., 2002
1B2 glycolipidNSP–COlfactory axon in olfactory bulb↑ cell–cell and cell–laminin adhesionMahanthappa et al., 1994
ActinNSP–PBrainNSJoubert et al., 1992
P–C (?)MOLT-4 T cellsPace et al., 1999
CA-125NSP–CHeLa cellsGal-1 export (?)Seelenmeyer et al., 2003
CD2/CD3DimericP–CJurkat T cellsMembrane redistribution, induction of cell deathPace et al., 1999; Walzel et al., 2000
CD4DimericP–CT cellNSPace et al., 1999; 2000
CD43DimericP–CT cellsMembrane redistribution, induction of cell deathPace et al., 1999; Nguyen et al., 2001; Lanteri et al., 2003
CD45DimericP–CT, B cellsMembrane redistribution, induction of cell deathPerillo et al., 1995; Fouillit et al., 2000; Symons et al., 2000; Fajka-Boja et al., 2002
CD7NSP–CT cellsInduction of cell deathPace et al., 1999, 2000
Carcino embryonic antigen (CEA, CD66e)NSP–CKM12 colon carcinoma cellsNSOhannesian et al., 1994
Cytochrome oxidase subunit IIINSP–P (?)HeLa cellsNSPark et al., 2001
FibronectinNSP–COvarian carcinoma, placenta↑ adhesionOzeki et al., 1995; Moiseeva et al., 2000; Moiseeva, Williams, and Samani, 2003; van den Brule et al., 2003
Genim-4 nuclear and (?) cytoplasmicNSP–PHeLa cellspreRNA splicing, RNA interferencePark et al., 2001; Hutvagner and Zamore, 2002
Glycoprotein 90K (MAC-2BP)NSP–CA375 melanoma cells↑ cell aggregationTinari et al., 2001
Glycosaminoglycan (chondroitin sulphate B, heparan sulfate)NSP–CVSMCModulation of ECM assembly, ↓ adhesionMoiseeva, Williams, Goodall et al., 2003
GM1 gangliosideDimericP–CSK-N-MC neurobastoma cells↓ growthKopitz et al., 1998, 2001; Andre et al., 2004
HBGp82NSP–CBrainNSChadli et al., 1997
H-rasDimericP–PHeLa, HEK293, Rat‐1, 293T cells↑ ras activation with selective activation of Raf-1/ERK pathwayPaz et al., 2001; Elad-Sfadia et al., 2002; Prior et al., 2003; Rotblat et al., 2004
LamininNSP–CMelanomas, myoblasts, ovarian carcinomas, Leydig cells, placenta↑ adhesionOzeki et al., 1995; Moiseeva et al., 2000; Moiseeva, Williams, Goodall et al., 2003; van den Brule et al., 2003; Martinez et al., 2004
LAMP-1 (CD107a), LAMP-2 (CD107b)NSP–COvarian, colon carcinomas↑ adhesionOhannesian et al., 1994; Woynarowska et al., 1994
MucinNSP–CEpithelial glycocalyces of gastric and intestinal mucosaNSWasano and Hirakawa, 1997
OsteospontinNSP–CVSMC↑ adhesionMoiseeva et al., 2000
Pre-B cell receptorNSP–CB cell lines↑ adhesion, cell diffrentiationGauthier et al., 2002
SUMO-3/SMT3BNSP–P (?)HeLa cellsNSPark et al., 2001
ThrombospondinNSP–CVSMC↑ adhesionMoiseeva et al., 2000
Thy-1NSP–CT cellsNSSymons et al., 2000
VitronectinNSP–CVSMCECM assemblyMoiseeva, Williams, and Samani, 2003

FAK, facal adhesion kinase; P–C, protein–carbohydrate interaction; P–P, protein–protein interaction; NS, not specified; VSMC, vascular smooth muscle cell; ↓, decreases and/or inhibits; ↑, increases and/or favors; ?, uncertain.

Protein–carbohydrate partnering

The lectin activity of Gal-1 relates to its carbohydrate-binding site (Figure 2). Sugar binding is enthalpically driven, and this supports the notion that van der Waals contacts and hydrogen bonds constitute the main forces driving and/or stabilizing complex formations (Lopez-Lucendo et al., 2004). The published dissociation constant of dGal-1 with various glycoproteins is about 5 µM (Symons et al., 2000). Although dGal-1 binds preferentially to glycoconjugates containing the ubiquitous disaccharide N-acetyllactosamine (Gal-β1–3/4 GlcNAc also known as LacNAcII or type 2 saccharide), its binding to individual lactosamine units is characterized by relatively low levels of affinity (Kd ∼ 50 µM) (Schwarz et al., 1998; Ahmad et al., 2004). It is the arrangement of lactosamine disaccharides in multiantennary repeating chains (up to three branches) that increases the binding avidity (Kd ∼ 4 µM) (Schwarz et al., 1998; Ahmad et al., 2004). In contrast, there is no increase in avidity when the recognition unit is repeated in a string (poly-N-lactosamine) (Ahmad et al., 2004). In polysaccharides, dGal-1 does not bind glycans that lack a terminal non-reducing unmodified N-acetyllactosamine (Di Virgilio et al., 1999; Stowell et al., 2004). Although terminal galactose residues are important for dGal-1 recognition, dGal-1 binds similarly to α3-sialylated (created by ST3Gal III sialyltransferase) and α2-fucosylated (created by fucosyltransferase) terminal N-acetyllactosamine, but not to α6‐sialylated or α3-fucosylated terminal N-acetyllactosamine (Amano et al., 2003; Leppanen et al., 2005). Whether extended or otherwise, free ligands in solution bind dGal-1 with a relatively low level of affinity (Leppanen et al., 2005); in contrast, the avidity of dGal-1 for extended glycans is enhanced when it is surface-bound as on cell surfaces or in extracellular matrix (ECM) (He and Baum, 2004).

In the ECM

Gal-1 binds to a number of ECM components in a dose-dependent and β-galactoside-dependent manner in the following order: laminin > cellular fibronectin > thrombospondin > plasma fibronectin > vitronectin > osteopontin (Moiseeva et al., 2000; Moiseeva, Williams, and Samani, 2003). Laminin and cellular fibronectin are glycoproteins which are highly N-glycosylated with bi- and tetra-antennary poly-N-lactosamines (Carsons et al., 1987; Fujiwara et al., 1988). Gal-1 is also involved in ECM assembly and remodeling: it inhibits the incorporation of vitronectin and chondroitin sulphate B into the ECM of vascular smooth muscle cells (Moiseeva, Williams, and Samani, 2003). The interaction of Gal-1 with vitronectin seems to depend on vitronectin conformation since it preferentially recognizes unfolded vitronectin multimers rather than inactive folded monomers (Moiseeva, Williams, and Samani, 2003).

Cell surface-binding partners

Various membrane glycoproteins have been identified as the binding partners of Gal-1 for the mediation of cell–cell or cell–ECM adhesion (Table II). We detail below some of the major cell surface-binding partners of Gal-1.

Integrins

The activity of integrin adhesion receptors is essential for normal cellular function and survival (Frisch and Ruoslahti, 1997; Stupack and Cheresch, 2002). N-glycosylations of β-integrins regulate β1 integrin functions by modulating their heterodimerization with α chains and ligand-binding activity (Gu and Taniguchi, 2004). Numerous variants of integrin glycoforms have been described in many normal and pathological cell types. Via its direct binding to β1 integrins (without cross-linking them) dGal-1 increases the amounts of partly activated β1 integrins, but does not induce dimerization with α subunits (Moiseeva, Williams, and Samani, 2003). In the case of vascular smooth muscle cells this interaction of Gal-1 with the α1β1 integrin has been reported both as transiently phosphorylating the focal adhesion kinase (FAK) and as modulating the attachment of cells and their spreading and migration on laminin, but not on cellular fibronectin (Moiseeva et al., 1999; Moiseeva, Williams, and Samani, 2003). Gal-1 is secreted during skeletal muscle differentiation and accumulates with laminin in the basement membrane surrounding each myofiber (Gu et al., 1994). The coincidence of Gal-1 secretion with the onset of myoblast differentiation and fusion and the transition in myoblast adhesion and mobility on laminin are regulated by the interaction of Gal-1 with laminin and the α7β1 integrin (Gu et al., 1994). As a consequence Gal-1 inhibits the association of the α7β1 integrin with laminin, and is thus able to prevent and dissociate the interaction of cells with laminin in a dose-dependent fashion (Gu et al., 1994). In contrast, Gal-1 does not affect the binding of cells on fibronectin via the same α7β1 integrin (Gu et al., 1994). Fischer et al. (2005) have recently shown on a panel of epithelial cell carcinomas that Gal-1-induced growth inhibition requires functional interactions with the α5β1 integrin.

As far as other integrins are concerned, Gal-1 from mouse macrophages has been found to specifically associate with the αMβ2 integrin (the complement receptor 3, CR3) (Avni et al., 1998).

CD2, CD3, CD7 CD43, CD45

A number of T-cell glycoproteins from MOLT-4 and Jurkat human T cells have been shown to be specific receptors for mammalian Gal-1 binding: CD45, CD43, CD7 (Pace et al., 1999; Walzel et al., 1999; Symons et al., 2000; Fajka-Boja et al., 2002). The functions of these receptors are detailed further in the review.

GM1 ganglioside

Gal-1 is a major receptor for the carbohydrate portion of the ganglioside GM1 exposed on the surface of human neuroblastoma cells (Kopitz et al., 1998; Andre et al., 2004). Cell confluence increases the surface presentation of dGal-1. Under these circumstances Gal-1 acts as a negative growth regulator of neuroblastoma cells, though without being pro-apoptotic (Kopitz et al., 1998, 2001).

Protein–protein partnering

The proteins identified so far that interact in a carbohydrate-independent manner with Gal-1 are not structurally related to each other and do not seem to share any common domains or motifs (Table II). The galectin sites that are involved in these interactions have not yet been established.

Gemin4

Gemin4 is found in the cytoplasm as well as in the nucleus of cells both as a member of the survival of motor neuron protein (SMN) complex and as the miRNP particle (microRNA [MiRNA] ribonucleoprotein [RNP]). The cytoplasmic SMN complex plays a role in the biogenesis of snRNPs in the cytoplasm before their entry into the nucleus (Paushkin et al., 2002). Nuclear SMN-containing complexes are thought to recycle and/or resupply snRNPs to the early (H/E) complexes in the spliceosome assembly pathway and so to be involved in the processes that direct pre-mRNA splicing (Paushkin et al., 2002). Thus, the findings that nuclear Gal-1 interacts with Gemin4 and is co-immunoprecipitated with nuclear SMN complexes (Park et al., 2001) offer mechanistic insights into its potential role in the splicing pathway (Vyakarnam et al., 1997) by involving the H/E complex as the locus of action of Gal-1 in the spliceosome assembly (Figure 3).

Ras

Gal-1 interacts in a lactose-independent manner with H-Ras-guanosine triphosphate (H-Ras-GTP) through its farnesyl cystein carboxymethylester (Paz et al., 2001; Rotblat et al., 2004) and so strengthens its membrane association (Paz et al., 2001). The binding of Gal-1 to Ras is one of the most interesting and potentially significant functions of Gal-1. We will therefore detail these functions in the following section.

The effect of Gal-1 on cell signaling pathways

Regulation of cell growth

While extracellular Gal-1 has no effect on the growth rates of naïve T cells (Endharti et al., 2005) or of astrocytic (Camby et al., 2002) or colon (Hittelet et al., 2003) tumor cell lines, Gal-1 is mitogenic for various types of normal or pathological murine and human cells, that is, murine Thy-1-negative spleen or lymph node cells (Symons et al., 2000), mammalian vascular cells (Sanford and Harris-Hooker, 1990; Moiseeva et al., 2000), and hepatic stellate cells (Maeda et al., 2003). Gal-1 inhibits the growth of other cell types such as neuroblastoma (Kopitz et al., 2001) and stromal bone marrow cells (Andersen et al., 2003). Interestingly enough, it has been reported that depending on the dose involved, Gal-1 causes the biphasic modulation of cell growth. While high doses (∼1 µM) of recombinant Gal-1 inhibit cell proliferation independently of Gal-1 sugar-binding activity, low doses (∼1 nM) of Gal-1 are mitogenic and are susceptible to inhibition by lactose (Adams et al., 1996; Vas et al., 2005). While the knock-down of Gal-1 expression in murine melanomas (Rubinstein, Alvarez et al., 2004) and human glioma cells (our unpublished data) does not affect their growth rate in vitro, it does decrease it in 9L rat gliosarcomas (Yamaoka et al., 2000). Furthermore, Gal-1 can also regulate cell cycle progression in human mammary tumor cells (Wells et al., 1999). The seemingly paradoxical positive and negative effects of Gal-1 on cell growth are highly dependent on cell type and cell activation status, and might also be influenced by the relative distribution of monomeric versus dimeric, or intracellular versus extracellular, forms.

Regulation of Cell Migration Processes

While cell migration is the net result of adhesion, motility, and invasion (Lefranc et al., 2005; Decaestecker et al., forthcoming), Gal-1 modifies each of these three cell migration-related processes.

Adhesion

Gal-1 has been shown to increase the adhesion of various normal and cancer cells to the ECM via the cross-linking of glycoproteins (integrins) exposed on the cell surfaces with carbohydrate moieties of ECM components such as laminin and fibronectin (Ellerhorst, Nguyen, Cooper, Lotan et al., 1999; Moiseeva et al., 1999; van den Brule et al., 2003). In addition, Gal-1 can also mediate homotypical cell interaction, so favoring the aggregation of human melanoma cells (Tinari et al., 2001) and heterotypical cell interactions such as the interaction between cancer and endothelial cells, which, in its turn, favors the dispersion of tumor cells (Clausse et al., 1999; Glinsky et al., 2000).

Motility

Gal-1 causes the increased motility of glioma cells and the reorganization of the actin cytoskeleton associated with an increased expression of RhoA, a protein that modulates actin polymerization and depolymerization (Camby et al., 2002) (Figure 3). Conversely, the knock-down of Gal-1 expression in glioma cells reduces motility and adhesiveness (Camby et al., 2002, 2005). Oxidized Gal-1 (see Gal-1 in pathological nervous systems) stimulates the migration of Schwann cells from both the proximal and the distal stumps of transected nerves and promotes axonal regeneration after peripheral nerve injury (Fukaya et al., 2003). In colon carcinomas a Gal-1-enriched ECM decreases colon carcinoma cell motility (Hittelet et al., 2003).

Invasion

Using a proteomic approach based on the comparison of highly and poorly invasive mammary carcinoma cell lines, Harvey et al. (2001) identified the membrane expression of Gal-1 as a signature of cell invasiveness.

Interaction between Gal-1 and Ras

Ras genes which are frequently mutated in human tumors promote malignant transformation (Paz et al., 2001). Ras transformation requires membrane anchorage, which is promoted by Ras farnesylcysteine carboxymethylester and a second signal. The overexpression of Gal-1 increases membrane-associated Ras, Ras-GTP, and active ERK results in cell transformations, which are blocked by dominant negative Ras (Paz et al., 2001). Gal-1 antisense RNA inhibits transformations by H-Ras and abolishes the membrane anchorage of green fluorescent protein (GFP)-H-Ras, but not of GFP-H-Ras wild-type, GFP-K-Ras, and GFP-N-Ras (Paz et al., 2001). Thus, H-Ras–Gal-1 interactions establish an essential link between two proteins associated with cell transformation and human malignancies that can be exploited to selectively target oncogenic Ras proteins. In fact, H-Ras-GTP recruits Gal-1 from the cytosol to the cell membrane with the resulting stabilization of H-Ras-GTP, the clustering of H-Ras-GTP and Gal-1 in non-raft microdomains (Prior et al., 2003), the subsequent binding to Raf-1 (but not to PI3Kinase), the activation of the ERK signaling pathway and, finally, increased cell transformation (Elad-Sfadia et al., 2002) (Figure 3). So, in addition to increasing and prolonging H-Ras activation, the Gal-1-H-Ras complex renders the activated molecule selective toward Raf-1, but not toward PI3K (Ashery et al., forthcoming) (Figure 3). Fischer et al. (2005) have observed that the antiproliferative potential of Gal-1 in a number of carcinoma cell lines requires functional interaction with the α5β1 integrin. Antiproliferative effects result from the inhibition of the Ras-MEK-ERK pathway and the consecutive transcriptional induction of p27, whose promoter contains two Sp1-binding sites crucial for Gal-1 responsiveness (Fischer et al., 2005). The inhibition of the Ras-MEK-ERK cascade by Gal-1 increases Sp1 transactivation and DNA binding due to the reduced threonine phosphorylation of Sp1. In addition, Gal-1 induces p21 transcription and selectively increases p27 protein stability, while the Gal-1-mediated accumulation of p27 and p21 inhibits cyclin-dependent kinase 2 activity, a process which ultimately results in G1 cell cycle arrest and growth inhibition (Fischer et al., 2005).

Rotblat et al. (2004) have identified a hydrophobic pocket in Gal-1 analogous to the Cdc42 geranylgeranyl-binding cavity in RhoGDI. This pocket possesses homologous isoprenoid-binding residues including the critical L11, whose RhoGDI L77 homologue changes dramatically on Cdc42 binding. By substituting L11A, Rotblat et al. (2004) obtained a dominant interfering Gal-1 that possesses a normal carbohydrate-binding ability but inhibits H-Ras GTP-loading and extracellular signal-regulated kinase activation, dislodges H-Ras from the cell membrane and attenuates H‐Ras fibroblast transformation and PC12-cell neurite outgrowth. Thus, whereas Gal-1 cooperates with Ras independently of carbohydrate binding, Gal-1 (L11A) inhibits it.

Gal-1 in embryonic and adult tissue development and differentiation

Menstrual cycle, early gestation and embryogenesis

Gal-1 expression has been reported in male and female gonads (Wollina et al., 1999; Timmons et al., 2002; Dettin et al., 2003), and exogenously added Gal-1 has an inhibitory effect both on the steroidogenic activity of Leydig cells in the testicles (Martinez et al., 2004) and on the granulosa cells in the ovary (Jeschke et al., 2004; Walzel et al., 2004).

In the uterus Gal-1 expression is restricted to the endometrium (Maquoi et al., 1997) and varies during the menstrual cycle and the early phases of gestation (von Wolff et al., 2005). The expression of Gal-1 increases significantly in the late secretory phase endometrium and in decidual tissue (Maquoi et al., 1997; von Wolff et al., 2005), and shows a specific pattern of expression in trophoblastic tissue (Maquoi et al., 1997; Vicovac et al., 1998).

During the first trimester of human embryogenesis Gal-1 is expressed in connective tissue, in smooth and striated muscles, and in some epithelia such as the skin, the gonads, the thyroid gland, and the kidneys (van den Brule et al., 1997; Savin et al., 2003; Hughes, 2004; von Wolff et al., 2005).

Differentiation of the myogenic lineage

During the course of myoblast differentiation intracellular Gal-1 is externalized as myoblasts fused into myotubes (Cooper and Barondes, 1990). The role of Gal-1 in the case of myoblast fusion may be explained by the fact that the adherence of the myoblast to the extracellular component laminin is disrupted in the presence of Gal-1 (Cooper et al., 1991) via the selective modulation by Gal-1 of the interaction between the α7β1 integrin and fibronectin and laminin (Gu et al., 1994). Although the exact role of Gal-1 in myogenesis remains to be seen, this galectin has been shown to induce non-committed myogenic cells in the dermis to express myogenic markers. It increases the terminal differentiation of committed myogenic cells and has a role to play in the development and regenerative ability of muscles (Cooper and Barondes, 1990; Cooper et al., 1991; Harrison and Wilson, 1992; Goldring et al., 2002). Gal-1 may thus be regarded as a potentially important tool in the treatment of cases of human muscular dystrophy (Goldring et al., 2002).

Differentiation of the hematopoietic lineage

Mesenchymal cells give rise to the stromal marrow environment that supports hematopoiesis. These cells constitute a wide range of differentiation potentials (e.g., adipocytes, osteoblasts, chondrocytes, lymphocytes, erythrocytes, macrophages) as well as a complex relationship with hematopoietic and endothelial cells. Numerous studies have demonstrated that Gal-1 may be a key element in the course of hematopoietic cell differentiation (Lutomski, Fouillit et al., 1997; Andersen et al., 2003; Silva et al., 2003; Wang et al., 2004; Vas et al., 2005). The K562 human leukemia cell line expresses Gal-1 in the cytosol, but upon treatment with erythropoietin these cells develop an erythroid phenotype that leads to the externalization of cytosolic Gal-1 (Lutomski, Fouillit et al., 1997). Similarly, Gal-1 is externalized during adipocyte differentiation (Wang et al., 2004) and is able to modulate osteoblastic differentiation (Andersen et al., 2003) as well as the proliferation and death of hematopoietic stem and progenitor cells (Vas et al., 2005).

Nerve structure development

Gal-1 is widely distributed in the central and peripheral nervous systems of rodents during their development. Although it has been shown that Gal-1 plays a number of important roles in the formation of the neural network of the olfactory bulb of mice (Puche et al., 1996), there are no reports on its role in other regions. Gal-1 homozygous null mutant (Gal-1–/–) mice are viable and can grow into adults without any obvious phenotypical abnormalities except for a deficiency in the olfactory network (Poirier and Robertson, 1993; Tenne-Brown et al., 1998) and a reduced thermal sensitivity (McGraw, Gaudet, Oschipok, Steeves et al., 2005). In these mice the neuronal subpopulation in the olfactory bulb, which normally expresses Gal-1, does not reach the appropriate targets in the olfactory glomeruli (Puche et al., 1996). During its development into adulthood, a rat’s sensory neurons from the dorsal root ganglion express Gal-1, as do some spinal motor neurons (Regan et al., 1986). The initial expression in the sensory neurons begins as they finish their final mitotic division and begin their growth toward their targets in the dorsal horn of the spinal cord. When Gal-1-expressing neurons reach their targets Gal-1 expression remains high, albeit at lower levels (Regan et al., 1986; Hynes et al., 1990; Sango et al., 2004). In addition to neurons, Gal-1 mRNAs are also detected in the non-neuronal cells such as the pia mater, the choroid plexus, and the pineal gland as well as in reactive astrocytic and Schwann cells (Akazawa et al., 2004; Sango et al., 2004; Egnaczyk et al., 2003).

Gal-1 and the immune system

Galectins in general, and Gal-1 in particular, are known to be deeply involved in the initiation, amplification, and resolution of inflammatory responses (Figure 4) (Almkvist and Karlsson, 2004).

Fig. 4.

Gal-1 and tumor immune privilege. The progression of the malignancy of a tumor includes the local dispersal of tumor cells into the surrounding normal tissue in addition to their long-range dispersal (metastasis). This progression is a multistep process that involves cell–cell and cell–ECM adhesion, invasion, migration, and angiogenesis. Gal-1 is involved in several of the steps of this malignant progression. Together with integrins, Gal-1 mediates tumor cell adhesion including adhesion to ECM proteins and homotypical cell adhesion. At the same time, Gal-1 can also inhibit adhesion, a phenomenon which could result in tumor cell detachment and migration. In addition, Gal-1 is involved in the mechanisms of tumor immune-escape. T cells use two main mechanisms to kill tumor cells, namely the death receptor and the granule exocytosis pathways, which involve the secretion of perforin and granzymes. Tumors can evade immune responses by secreting immunosuppressive cytokines and soluble inhibitory factors, including Gal-1. Gal-1 contributes to immune evasion by inducing apoptosis in effector T cells (Rubinstein, Alvarez et al., 2004; Liu and Rabinovich, 2005).

T-cell homeostasis and survival

A growing body of evidence indicates that Gal-1 functions as a homeostatic agent by modulating innate and adaptative immune responses. Gal-1 induces the inhibition of cell growth and cell-cycle arrest (Blaser et al., 1998; Rabinovich, Ramhorst et al., 2002) and promotes the apoptosis of activated, but not resting, immune cells (Perillo et al., 1995; Rabinovich et al., 1998; Chung et al., 2000; He and Baum, 2004). This said, resting T cells are sensitized to CD95/Fas-mediated cell death by Gal-1 (Matarrese et al., 2005). Furthermore, it has been shown that the Gal-1 expressed by thymic epithelial cells promotes the apoptosis of immature cortical thymocytes in vitro (Perillo et al., 1997), so suggesting a potential role for this protein in the processes of positive and/or negative selection within the thymic microenvironment. Gal-1 also suppresses the secretion of the pro-inflammatory cytokine interleukin-2 (IL-2) (Rabinovich, Ariel et al., 1999) and favors the secretion of the anti-inflammatory cytokine IL-10 (van der Leij et al., 2004) (Figure 4). All of these activities have been demonstrated by adding a relatively high concentration (µM range) of exogenous Gal-1 to T cells in vitro. In this context Bättig et al. (2004) have shown that the irreversibly dimeric form of Gal-1 is a dramatically more potent inducer of apoptosis in T cells than wild-type Gal-1. One concern regarding the proapoptotic activity of Gal-1 is whether high levels of soluble protein can be achieved in vivo. Recent evidence indicates that the amount of Gal-1 secreted by different cell types in the ECM is sufficient to kill T cells (Perillo et al., 1995; Chung et al., 2000; He and Baum, 2004). The effects of Gal-1 on immune and inflammatory cells are likely to be due to the binding and cross-linking of cell-surface glycoproteins on these cells (Galvan et al., 2000) (Figure 4). As a bivalent dimer, Gal-1 binds to the glycoproteins (including CD2, CD3, CD7, CD43, and CD45) on the cell surface of T cells in a carbohydrate-dependent manner (Pace et al., 1999). The regulated expression of glycosyltransferases—leading to the creation of N-acetyllactosamine ligands—during development and activation may determine T-cell susceptibility to Gal-1-induced cell death (Galvan et al., 2000; Amano et al., 2003; Carlow et al., 2003). Sezary cells—the malignant T cells in cutaneous T-cell lymphomas (the Sezary syndrome or mycosis fungoides)—resist a variety of apoptosis-inducing agents including Gal-1, because of the loss of CD7 expression and altered cellular glycosylations (Rappl et al., 2002; Roberts et al., 2003).

Thus, a number of T-cell glycoproteins from human MOLT-4 and Jurkat T cells have been shown to be specific receptors for mammalian Gal-1 binding and to be involved in Gal-1-mediated T-cell death: CD45, CD43, CD7 (Pace et al., 1999; Walzel et al., 1999; Symons et al., 2000; Fajka-Boja et al., 2002). However, although the deletion mutants of the glycoproteins confirm their importance in the apoptotic response to Gal-1 (Pace et al., 1999, 2000; Perillo et al., 1995), the role of CD45 in T-cell apoptosis mediated by Gal-1 remains controversial since Gal-1 induces apoptosis in CD45-deficient T cells (Walzel et al., 1999; Fajka-Boja et al., 2002). As for CD7, it seems that only specific spliced isoforms or glycoforms of CD45 may be important in signaling Gal-1-induced cell death (Nguyen et al., 2001; Xu and Weiss, 2002; Amano et al., 2003; Lanteri et al., 2003).

The signal transduction events that lead to galectin-induced cell death in activated T cells involve several intracellular mediators including the induction of specific transcription factors (i.e., NFAT, AP-1), the activation of the Lck/ZAP-70/MAPK signaling pathway, the modulation of Bcl-2 protein production, the depolarization of the mitochondrial membrane potential and cytochrome c release, the activation of caspases and the participation of the ceramide pathway (Rabinovich, Alonso et al., 2000; Walzel et al., 2000; Hahn et al., 2004; Ion et al., 2005; Matarrese et al., 2005) (Figure 4). However, a recent study has shown that Gal-1-induced apoptosis in human T leukemia MOLT-4 cells deficient in Fas-induced cell death is not dependent on the activation of caspase-3 or on cytochrome c release—two hallmarks of apoptosis—but involves the rapid nuclear translocation of EndoG from mitochondria (Hahn et al., 2004), so implying that Gal-1-induced cell death might also relate to one of the other types of cell death (Broker et al., 2005). Furthermore, recent evidence also indicates that whereas dGal-1 can induce the exposure of phosphatidylserine (an early apoptotic marker involved in the phagocytosis of apoptotic cells) on the plasma membrane of human T leukemia MOLT-4 cells, this does not result in cell death on activated neutrophils and on the promyelocytic cell line, but prepares the cells for phagocytic removal (Dias-Baruffi et al., 2003). At low concentrations (the nM range) Gal-1 has been shown in vitro to inhibit T‐cell adhesion to ECM and to modulate the tumor necrosis factor alpha (TNFα) as well as the interferon gamma (IFNγ) secretion from activated T cells (Allione et al., 1998) (Figure 4). In addition, in vivo studies on experimental autoimmunity models have revealed the ability of Gal-1 to skew the balance toward a T2-type cytokine response by reducing the levels of IFNγ, TNFα, IL-2, and IL-12 and increasing the level of IL-5 secretion (Santucci et al., 2000, 2003; Baum et al., 2003) (Figure 4).

T-cell immune disorders and chronic inflammation

In vivo, Gal-1 has powerful immunoregulatory effects through its ability to inhibit T-cell effector functions (van der Leij et al., 2004; Figure 4). Gal-1 treatment has resulted in improvements and even in cases of prevention in a number of experimental models of autoimmune diseases (Table III). The in vivo administration of Gal-1 prevents the development of chronic inflammation and impairs the ongoing disease in experimental models of autoimmune encephalomyelitis (EAE) (Offner et al., 1990), arthritis (Rabinovich, Daly et al., 1999), colitis (Santucci et al., 2003), hepatitis (Santucci et al., 2000), and chronic pancreatitis (Wang et al., 2000). The ability of Gal-1 to suppress the allogenic T-cell response through apoptotic and non-apoptotic mechanisms suggests its potential use for immunosuppression in organ transplantation and graft versus host disease (GVHD) (Baum et al., 2003).

Table III.

Overview of animal models that have undergone Gal-1 treatment

Human diseaseAnimal modelTreatmentResultsReferences
Inflammation-related diseasesCrohn’s diseaseTrinitrobenzene sulphonic acid-induced colitis in BALBc micerhGal-1 (1 mg/kg), prophylactic or therapeuticClinical and histological improvementSantucci et al., 2003
Multiple sclerosisAutoimmune encephalomyelitis (EAE) in Lewis ratsrhGal-1 (250 µg i.v. for 10 days, start day –3, or i.v. daily for 12 days, start day 0)63% protection against disease (prophylactic)Offner et al., 1990
90% no clinical signs (therapeutical)
Myasthenia gravisExperimental autoimmune myasthenia gravis in New Zealand rabbitsGal-1 (electrolectin) (prophylactic and therapeutical protocols)Prophylactic and therapeutical effectsLevi et al., 1983
Rheumatoid arthritisCollagen-induced arthritis in DBA/1 miceGene therapy: fibroblast secreting moGal-1 or daily i.p. of rhGal-1Therapeutic effects in both protocols, skewing Th1 ≥ Th2Rabinovich, Daly et al., 1999
Nephritis (Goodpasture’s syndrome)Nephrotoxic nephritis in Wistar Kyoto ratsrmGal-1 (1 mg/kg, every second day, 2 weeks)Therapeutic effectsTsuchiyama et al., 2000
T-cell mediated hepatitisCon A-induced hepatitis in BALBc micerhGal-1 (i.v. at time of induction)Prevention of liver injurySantucci et al., 2000
GVHDBM Tx in micerhGal-1 (i.p., 3×/week)Reduced mortalityBaum et al., 2003
CancersHuman glioblastomasIntracranial xenografts in nude miceTransfection of anti Gal-1 antisense oligonucleotides prior to graftReduced mortalityCamby et al., 2002
MelanomasMice with subcutaneous melanoma graftsTransfection of anti Gal-1 antisense oligonucleotides prior to graftReduced mortality, tumor immune-escapeRubinstein, Alvarez et al., 2004
Neuro-regenerationALSH46R SOD1 transgenic micerGal-1/oxidized (i.m., 0.25 µg/g/week)Delay of the onset of the disease, prolonged lifespan, and improved motor functionChang-Hong et al., 2005
Peripheral nerve injuryRats with surgically transected sciatic nervesRhGal-1/oxidized (5 µg/mL by osmotic pump, 2.5 µl/h at the site of surgery)Functional recoveryKadoya and Horie, 2005
Human diseaseAnimal modelTreatmentResultsReferences
Inflammation-related diseasesCrohn’s diseaseTrinitrobenzene sulphonic acid-induced colitis in BALBc micerhGal-1 (1 mg/kg), prophylactic or therapeuticClinical and histological improvementSantucci et al., 2003
Multiple sclerosisAutoimmune encephalomyelitis (EAE) in Lewis ratsrhGal-1 (250 µg i.v. for 10 days, start day –3, or i.v. daily for 12 days, start day 0)63% protection against disease (prophylactic)Offner et al., 1990
90% no clinical signs (therapeutical)
Myasthenia gravisExperimental autoimmune myasthenia gravis in New Zealand rabbitsGal-1 (electrolectin) (prophylactic and therapeutical protocols)Prophylactic and therapeutical effectsLevi et al., 1983
Rheumatoid arthritisCollagen-induced arthritis in DBA/1 miceGene therapy: fibroblast secreting moGal-1 or daily i.p. of rhGal-1Therapeutic effects in both protocols, skewing Th1 ≥ Th2Rabinovich, Daly et al., 1999
Nephritis (Goodpasture’s syndrome)Nephrotoxic nephritis in Wistar Kyoto ratsrmGal-1 (1 mg/kg, every second day, 2 weeks)Therapeutic effectsTsuchiyama et al., 2000
T-cell mediated hepatitisCon A-induced hepatitis in BALBc micerhGal-1 (i.v. at time of induction)Prevention of liver injurySantucci et al., 2000
GVHDBM Tx in micerhGal-1 (i.p., 3×/week)Reduced mortalityBaum et al., 2003
CancersHuman glioblastomasIntracranial xenografts in nude miceTransfection of anti Gal-1 antisense oligonucleotides prior to graftReduced mortalityCamby et al., 2002
MelanomasMice with subcutaneous melanoma graftsTransfection of anti Gal-1 antisense oligonucleotides prior to graftReduced mortality, tumor immune-escapeRubinstein, Alvarez et al., 2004
Neuro-regenerationALSH46R SOD1 transgenic micerGal-1/oxidized (i.m., 0.25 µg/g/week)Delay of the onset of the disease, prolonged lifespan, and improved motor functionChang-Hong et al., 2005
Peripheral nerve injuryRats with surgically transected sciatic nervesRhGal-1/oxidized (5 µg/mL by osmotic pump, 2.5 µl/h at the site of surgery)Functional recoveryKadoya and Horie, 2005

i.m., intramuscular injection; i.p., intraperitoneal injection; i.v., intravascular injection; moGal-1, mouse galectin-1; rhGal-1, recombinant human galectin-1.

Table III.

Overview of animal models that have undergone Gal-1 treatment

Human diseaseAnimal modelTreatmentResultsReferences
Inflammation-related diseasesCrohn’s diseaseTrinitrobenzene sulphonic acid-induced colitis in BALBc micerhGal-1 (1 mg/kg), prophylactic or therapeuticClinical and histological improvementSantucci et al., 2003
Multiple sclerosisAutoimmune encephalomyelitis (EAE) in Lewis ratsrhGal-1 (250 µg i.v. for 10 days, start day –3, or i.v. daily for 12 days, start day 0)63% protection against disease (prophylactic)Offner et al., 1990
90% no clinical signs (therapeutical)
Myasthenia gravisExperimental autoimmune myasthenia gravis in New Zealand rabbitsGal-1 (electrolectin) (prophylactic and therapeutical protocols)Prophylactic and therapeutical effectsLevi et al., 1983
Rheumatoid arthritisCollagen-induced arthritis in DBA/1 miceGene therapy: fibroblast secreting moGal-1 or daily i.p. of rhGal-1Therapeutic effects in both protocols, skewing Th1 ≥ Th2Rabinovich, Daly et al., 1999
Nephritis (Goodpasture’s syndrome)Nephrotoxic nephritis in Wistar Kyoto ratsrmGal-1 (1 mg/kg, every second day, 2 weeks)Therapeutic effectsTsuchiyama et al., 2000
T-cell mediated hepatitisCon A-induced hepatitis in BALBc micerhGal-1 (i.v. at time of induction)Prevention of liver injurySantucci et al., 2000
GVHDBM Tx in micerhGal-1 (i.p., 3×/week)Reduced mortalityBaum et al., 2003
CancersHuman glioblastomasIntracranial xenografts in nude miceTransfection of anti Gal-1 antisense oligonucleotides prior to graftReduced mortalityCamby et al., 2002
MelanomasMice with subcutaneous melanoma graftsTransfection of anti Gal-1 antisense oligonucleotides prior to graftReduced mortality, tumor immune-escapeRubinstein, Alvarez et al., 2004
Neuro-regenerationALSH46R SOD1 transgenic micerGal-1/oxidized (i.m., 0.25 µg/g/week)Delay of the onset of the disease, prolonged lifespan, and improved motor functionChang-Hong et al., 2005
Peripheral nerve injuryRats with surgically transected sciatic nervesRhGal-1/oxidized (5 µg/mL by osmotic pump, 2.5 µl/h at the site of surgery)Functional recoveryKadoya and Horie, 2005
Human diseaseAnimal modelTreatmentResultsReferences
Inflammation-related diseasesCrohn’s diseaseTrinitrobenzene sulphonic acid-induced colitis in BALBc micerhGal-1 (1 mg/kg), prophylactic or therapeuticClinical and histological improvementSantucci et al., 2003
Multiple sclerosisAutoimmune encephalomyelitis (EAE) in Lewis ratsrhGal-1 (250 µg i.v. for 10 days, start day –3, or i.v. daily for 12 days, start day 0)63% protection against disease (prophylactic)Offner et al., 1990
90% no clinical signs (therapeutical)
Myasthenia gravisExperimental autoimmune myasthenia gravis in New Zealand rabbitsGal-1 (electrolectin) (prophylactic and therapeutical protocols)Prophylactic and therapeutical effectsLevi et al., 1983
Rheumatoid arthritisCollagen-induced arthritis in DBA/1 miceGene therapy: fibroblast secreting moGal-1 or daily i.p. of rhGal-1Therapeutic effects in both protocols, skewing Th1 ≥ Th2Rabinovich, Daly et al., 1999
Nephritis (Goodpasture’s syndrome)Nephrotoxic nephritis in Wistar Kyoto ratsrmGal-1 (1 mg/kg, every second day, 2 weeks)Therapeutic effectsTsuchiyama et al., 2000
T-cell mediated hepatitisCon A-induced hepatitis in BALBc micerhGal-1 (i.v. at time of induction)Prevention of liver injurySantucci et al., 2000
GVHDBM Tx in micerhGal-1 (i.p., 3×/week)Reduced mortalityBaum et al., 2003
CancersHuman glioblastomasIntracranial xenografts in nude miceTransfection of anti Gal-1 antisense oligonucleotides prior to graftReduced mortalityCamby et al., 2002
MelanomasMice with subcutaneous melanoma graftsTransfection of anti Gal-1 antisense oligonucleotides prior to graftReduced mortality, tumor immune-escapeRubinstein, Alvarez et al., 2004
Neuro-regenerationALSH46R SOD1 transgenic micerGal-1/oxidized (i.m., 0.25 µg/g/week)Delay of the onset of the disease, prolonged lifespan, and improved motor functionChang-Hong et al., 2005
Peripheral nerve injuryRats with surgically transected sciatic nervesRhGal-1/oxidized (5 µg/mL by osmotic pump, 2.5 µl/h at the site of surgery)Functional recoveryKadoya and Horie, 2005

i.m., intramuscular injection; i.p., intraperitoneal injection; i.v., intravascular injection; moGal-1, mouse galectin-1; rhGal-1, recombinant human galectin-1.

Acute inflammation and allergy

In addition to its role in adaptative immune responses and chronic inflammation, Gal-1 also participates in innate immunity and acute and allergic inflammation (Liu, 2000). In phospholipase A2 (PLA2)-induced hind paw edema in rats the transmigration of both neutrophils and mast cells into the tissue is reduced in the presence of Gal-1 (Rubinstein, Ilarregui et al., 2004), which is also responsible for the inhibition of the release of arachidonic acid from lipopolysaccharide (LPS)-stimulated macrophages, mast-cell degranulation, and eosinophil migration (Rabinovich, Sotomayor et al., 2000; Delbrouck et al., 2002; La et al., 2003). This suggests that signals generated by Gal-1 binding inhibit rather than promote the migration of inflammatory cells.

Gal-1 is also able to induce an oxidative burst in neutrophils that have extravasated into tissue, but not in the case of peripheral blood neutrophils (Almkvist et al., 2002). This enhancement of cellular activity in exudated neutrophils is referred to as the neutrophil priming that might occur through the interaction of Gal-1 with αMβ2 integrins expressed at the neutrophil cell surface (Almkvist et al., 2002), as has been described with respect to macrophages (Avni et al., 1998).

Host–pathogen (bacteria–virus–parasites) interactions

Taking recombinant Gal-1 as a basis, it has been shown that this galectin influences the ability of macrophages to control intracellular infections by inhibiting microbicidal activity, by promoting parasite replication, or by inducing host-cell apoptosis (Zuniga et al., 2001). While a biphasic modulation has been reported in Trypanosoma cruzi replication and cell viability, low concentrations (3 nM) of Gal‐1 increase parasite replication and do not affect macrophage survival, higher concentrations (300 nM) are able to condemn cells to apoptosis and to inhibit parasite replication. The expression of Gal-1 is markedly upregulated after parasite or virus infection (Giordanengo et al., 2001; Zuniga et al., 2001; Lim et al., 2003). It has been shown on the basis of experiments using exogenously added recombinant protein that Gal-1 acts as a soluble host factor that promotes HIV-1 infectivity though the stabilization of virus attachment to host cells (Ouellet et al., 2005), and that the altered T-cell surface glycosylations in HIV-1 infection results in an increased susceptibility to Gal-1-induced cell death (Lanteri et al., 2003). In contrast, in the case of the Nipah virus infection (responsible for severe, often fatal, febrile encephalitis) of endothelial cells, dGal-1 inhibits the cell fusion of the envelope glycoproteins of the Nipah virus with the host cells and favors the secretion of proinflammatory cytokines by dendritic cells (Levroney et al., 2005).

Gal-1 involvement in tumor progression and tumor immune-escape

From all the studies reported in the literature and summarized in Table IV it is reasonable to assume that Gal-1 expression or overexpression in a tumor or the tissue surrounding a tumor (stroma) must be considered as a sign of the tumor’s malignant progression and, consequently, of a poor prognosis for patients. This prognosis is often related to tumor immune-escape, to the long-range dissemination of tumoral cells (metastasis), or to their presence in the surrounding normal tissue, as is discussed below and illustrated in Figure 4.

Table IV.

Gal-1 expression in human tumors

Histological typesExpression in tumors as compared to normal tissuesIs Gal-1 a diagnostic marker?Is Gal-1 a prognostic marker?Does Gal-1 modify cell proliferation and/or cell migration?
Colon carcinomas↑ in stroma and in epithelial tissues (Sanjuan et al., 1997; Hittelet et al., 2003; Nagy et al., 2003)Yes (Nagy et al., 2003)Yes (Hittelet et al., 2003; Horiguchi et al., 2003)
Pancreatic ductal adenocarcinomas↑ (Grutzmann et al. 2004; Shen et al., 2004)Yes (Berberat et al., 2001; Fitzner et al., 2005)
Intrahepatic cholangiocarcinomas↑ in stroma and in epithelial tissues (Shimonishi et al., 2001)
Renal cell carcinomas↑ or ↓ depending on histological grades (Francois et al., 1999; Saussez et al., 2005)
Bladder transitional-cell carcinomas↑ (Cindolo et al., 1999)Yes (Cindolo et al., 1999)
Prostate cancers↑ in stromal tissues (van den Brule et al., 2001)Yes (van den Brule et al., 2001)
Uterine adenocarcinomas↑ (van den Brule et al., 1996)
Choriocarcinomas↑ (Bozic et al., 2004)
Human uterine smooth muscle tumorsNo modifications (Schwarz et al., 1999)
Gliomas↑ (Gunnersen et al., 2000; Yamaoka et al., 2000;     Camby et al., 2001, 2002; Rorive et al., 2001)Yes (Camby et al., 2001, 2002)Yes (Camby et al., 2001, 2002; Rorive et al., 2001)
Nonsmall-cell lung cancers↑ (Szoke et al., 2005)Yes (Gabius et al., 2002)
HNSCCs↑ or ↓ depending on histological types (Gillenwater et al., 1996; Choufani et al., 1999; He et al., 2004)Yes (Le et al., 2005; Saussez et al., forthcoming)
Histological typesExpression in tumors as compared to normal tissuesIs Gal-1 a diagnostic marker?Is Gal-1 a prognostic marker?Does Gal-1 modify cell proliferation and/or cell migration?
Colon carcinomas↑ in stroma and in epithelial tissues (Sanjuan et al., 1997; Hittelet et al., 2003; Nagy et al., 2003)Yes (Nagy et al., 2003)Yes (Hittelet et al., 2003; Horiguchi et al., 2003)
Pancreatic ductal adenocarcinomas↑ (Grutzmann et al. 2004; Shen et al., 2004)Yes (Berberat et al., 2001; Fitzner et al., 2005)
Intrahepatic cholangiocarcinomas↑ in stroma and in epithelial tissues (Shimonishi et al., 2001)
Renal cell carcinomas↑ or ↓ depending on histological grades (Francois et al., 1999; Saussez et al., 2005)
Bladder transitional-cell carcinomas↑ (Cindolo et al., 1999)Yes (Cindolo et al., 1999)
Prostate cancers↑ in stromal tissues (van den Brule et al., 2001)Yes (van den Brule et al., 2001)
Uterine adenocarcinomas↑ (van den Brule et al., 1996)
Choriocarcinomas↑ (Bozic et al., 2004)
Human uterine smooth muscle tumorsNo modifications (Schwarz et al., 1999)
Gliomas↑ (Gunnersen et al., 2000; Yamaoka et al., 2000;     Camby et al., 2001, 2002; Rorive et al., 2001)Yes (Camby et al., 2001, 2002)Yes (Camby et al., 2001, 2002; Rorive et al., 2001)
Nonsmall-cell lung cancers↑ (Szoke et al., 2005)Yes (Gabius et al., 2002)
HNSCCs↑ or ↓ depending on histological types (Gillenwater et al., 1996; Choufani et al., 1999; He et al., 2004)Yes (Le et al., 2005; Saussez et al., forthcoming)

↓, decreased Gal-1 expression in tumor tissue as compared to normal tissues; ↑, increased Gal-1 expression in tumor tissue as compared to normal tissues.

Table IV.

Gal-1 expression in human tumors

Histological typesExpression in tumors as compared to normal tissuesIs Gal-1 a diagnostic marker?Is Gal-1 a prognostic marker?Does Gal-1 modify cell proliferation and/or cell migration?
Colon carcinomas↑ in stroma and in epithelial tissues (Sanjuan et al., 1997; Hittelet et al., 2003; Nagy et al., 2003)Yes (Nagy et al., 2003)Yes (Hittelet et al., 2003; Horiguchi et al., 2003)
Pancreatic ductal adenocarcinomas↑ (Grutzmann et al. 2004; Shen et al., 2004)Yes (Berberat et al., 2001; Fitzner et al., 2005)
Intrahepatic cholangiocarcinomas↑ in stroma and in epithelial tissues (Shimonishi et al., 2001)
Renal cell carcinomas↑ or ↓ depending on histological grades (Francois et al., 1999; Saussez et al., 2005)
Bladder transitional-cell carcinomas↑ (Cindolo et al., 1999)Yes (Cindolo et al., 1999)
Prostate cancers↑ in stromal tissues (van den Brule et al., 2001)Yes (van den Brule et al., 2001)
Uterine adenocarcinomas↑ (van den Brule et al., 1996)
Choriocarcinomas↑ (Bozic et al., 2004)
Human uterine smooth muscle tumorsNo modifications (Schwarz et al., 1999)
Gliomas↑ (Gunnersen et al., 2000; Yamaoka et al., 2000;     Camby et al., 2001, 2002; Rorive et al., 2001)Yes (Camby et al., 2001, 2002)Yes (Camby et al., 2001, 2002; Rorive et al., 2001)
Nonsmall-cell lung cancers↑ (Szoke et al., 2005)Yes (Gabius et al., 2002)
HNSCCs↑ or ↓ depending on histological types (Gillenwater et al., 1996; Choufani et al., 1999; He et al., 2004)Yes (Le et al., 2005; Saussez et al., forthcoming)
Histological typesExpression in tumors as compared to normal tissuesIs Gal-1 a diagnostic marker?Is Gal-1 a prognostic marker?Does Gal-1 modify cell proliferation and/or cell migration?
Colon carcinomas↑ in stroma and in epithelial tissues (Sanjuan et al., 1997; Hittelet et al., 2003; Nagy et al., 2003)Yes (Nagy et al., 2003)Yes (Hittelet et al., 2003; Horiguchi et al., 2003)
Pancreatic ductal adenocarcinomas↑ (Grutzmann et al. 2004; Shen et al., 2004)Yes (Berberat et al., 2001; Fitzner et al., 2005)
Intrahepatic cholangiocarcinomas↑ in stroma and in epithelial tissues (Shimonishi et al., 2001)
Renal cell carcinomas↑ or ↓ depending on histological grades (Francois et al., 1999; Saussez et al., 2005)
Bladder transitional-cell carcinomas↑ (Cindolo et al., 1999)Yes (Cindolo et al., 1999)
Prostate cancers↑ in stromal tissues (van den Brule et al., 2001)Yes (van den Brule et al., 2001)
Uterine adenocarcinomas↑ (van den Brule et al., 1996)
Choriocarcinomas↑ (Bozic et al., 2004)
Human uterine smooth muscle tumorsNo modifications (Schwarz et al., 1999)
Gliomas↑ (Gunnersen et al., 2000; Yamaoka et al., 2000;     Camby et al., 2001, 2002; Rorive et al., 2001)Yes (Camby et al., 2001, 2002)Yes (Camby et al., 2001, 2002; Rorive et al., 2001)
Nonsmall-cell lung cancers↑ (Szoke et al., 2005)Yes (Gabius et al., 2002)
HNSCCs↑ or ↓ depending on histological types (Gillenwater et al., 1996; Choufani et al., 1999; He et al., 2004)Yes (Le et al., 2005; Saussez et al., forthcoming)

↓, decreased Gal-1 expression in tumor tissue as compared to normal tissues; ↑, increased Gal-1 expression in tumor tissue as compared to normal tissues.

Gal-1 could be involved in tumor angiogenesis because both vascular smooth muscle and endothelial cells express it (Moiseeva et al., 2000; Moiseeva, Williams, and Samani, 2003). Although the vessel walls of normal lymphoid tissues do not express Gal-1, the blood vessel walls of lymphomas do so in relation to their vascular density (D’Haene et al., 2005).

A number of mechanisms have been described that potentially contribute to tumor cell evasion from an anti-tumoral immune response (Zou, 2005). Together with the abundance of pro-apoptotic Gal-1 in privileged immune sites such as the placenta (Hirabayashi et al., 1989), the brain (Joubert et al., 1989), and the reproductive organs (Wollina et al., 1999) the fact that a number of studies have highlighted the expression of Gal-1 in the stromal tissue around tumors (Gillenwater et al., 1996; Sanjuan et al., 1997; Berberat et al., 2001; Shimonishi et al., 2001; van den Brule et al., 2001, 2003) or in the endothelial cells from capillaries infiltrating them rather than in those in the adjacent non-tumoral stroma (Clausse et al., 1999) suggest that Gal-1 might trigger the death of infiltrating T cells and protect these sites from the tissue damage induced by T-cell-derived proinflammatory cytokines (Figure 4). Le et al. (2005) have recently demonstrated a significant relation in head and neck squamous cell carcinomas (HNSCCs) between Gal-1 expression and the presence of both hypoxia markers and an inverse correlation with T-cell infiltration, a fact which suggests that hypoxia can affect malignant progression by regulating the secretion by tumor cells of proteins (like Gal-1) that modulate immune privilege. The immunomodulatory effects of Gal-1 and the correlation between Gal-1 expression in cancer cells and their aggressiveness (as described in the previous section) suggest the hypothesis that tumor cells may impair T-cell effector functions through the secretion of Gal-1, and that this mechanism may contribute toward tilting the balance in favor of an immunosuppressive environment at a tumor site (Figure 4). The blockade of the biological activity of Gal-1 in melanoma tissue results in a reduced tumor mass and stimulates the in vivo generation of a tumor-specific T-cell response (Rubinstein, Alvarez et al., 2004). These observations support the idea that Gal-1 may contribute to the immune privilege of tumors by modulating the survival or polarization of effector T cells (Figure 4).

Gal-1 in pathological nervous systems

Nerve regeneration

Gal-1 in its oxidized form—a form that lacks lectin activity—promotes neurite outgrowth (Outenreath and Jones, 1992) and enhances axonal regeneration in peripheral (Horie et al., 1999; Inagaki et al., 2000; Fukaya et al., 2003; Kadoya et al., 2005) and central (McGraw, McPhail et al., 2004; Rubinstein, Ilarregui et al., 2004; McGraw, Gaudet, Oschipok, Kadoya et al., 2005) nerves even at relatively low concentrations (picoM range) (Horie and Kadoya, 2000). The marked axonal regeneration-promoting activity of oxidized Gal-1 is likely to be paracrine (Figure 5). Indeed, Gal-1 is expressed in dorsal root ganglion neurons and motor neurons, with immunoreactivity restricted to the neuronal cell bodies, the axons, and the Schwann cells of adult rodents (Regan et al., 1986; Hynes et al., 1990; Horie et al., 1999; Fukaya et al., 2003) (Figure 5). After axonal injury, cytosolic reduced Gal-1 is likely to be externalized from growing axons and reactive Schwann cells to an extracellular space where some of the molecules may be converted into an oxidized form and may enhance axonal regeneration (Horie and Kadoya, 2000; McGraw, McPhail et al., 2004; Rubinstein, Ilarregui et al., 2004; Miura et al., 2004; Sango et al., 2004) (Figure 5). Miura et al. (2004) have recently identified a novel, naturally occurring, N-terminally processed form of Gal-1 that lacks the six amino-terminal residues of full length Gal-1. This isoform of Gal-1, which is monomeric under both reducing and oxidizing conditions, promotes axonal regeneration (Miura et al., 2004). Since oxidized Gal-1-induced neurite outgrowth is not observed on isolated neurons (Horie et al., 1999), the secreted Gal-1 probably influences the non-neuronal cells surrounding the axons, including the Schwann cells (Fukaya et al., 2003), and in so doing recruits macrophages, fibroblasts, and perineuronal cells (Horie et al., 2004) (Figure 5). In this respect, macrophages are potential candidates since they secrete an axonal regeneration-promoting factor when stimulated by oxidized Gal-1 (Horie et al., 2004). A preclinical study using rats with surgically transected sciatic nerves has recently shown that the administration by an osmotic pump of oxidized Gal-1 at the site of surgery restores nerve function (Kadoya et al., 2005).

Fig. 5.

Gal-1 and neuroregeneration processes. After axonal injury, cytosolic-reduced Gal-1 is externalized from growing axons and reactive Schwann cells to the extracellular space, where some of the molecules are converted into an oxidized form (Horie and Kadoya, 2000; McGraw, McPhail et al., 2004; Rubinstein, Ilarregui et al., 2004; Miura et al., 2004; Sango et al., 2004). Upon stimulation by oxidized Gal-1, macrophages secrete an axonal regeneration-promoting factor and enhance axonal regeneration (Horie et al., 2004). Reduced Gal-1 induces astrocyte differentiation in a carbohydrate-dependent manner. The differentiated astrocytes then greatly enhance their production of the BDNF that, in turn, plays an important role in the survival, differentiation, and synaptic plasticity of neurons (Sasaki et al., 2004).

Neural diseases

As mentioned above, Gal-1 is an endogenously expressed protein that is important in the embryonic development of primary sensory neurons and their synaptic connections in the spinal cord (Puche et al., 1996; Tenne-Brown et al., 1998). Various reports suggest a relation between Gal-1 expression (or altered expression) and neurological diseases. Gal-1 expression by neuronal and glial cells is closely correlated with regenerative success after injury (Wada et al., 2003; McGraw, Gaudet, Oschipok, Kadoya et al., 2005), and the level of autoantibodies to Gal-1 is significantly higher in patients with neurological disorders than in healthy controls (Lutomski, Joubert-Caron et al., 1997). In cases of neurodegenerative amyotrophic lateral sclerosis (ALS) diseases Gal-1 accumulates in the neurofilamentous lesions (Wada et al., 2003). In a recent report, Chang-Hong et al. (2005) show the neuroprotective effect of oxidized Gal-1 on a transgenic murine ALS model: the administration of oxidized Gal-1 to the mice delayed the onset of their disease, prolonged their life spans, and improved their motor functions.

In the rat brain hippocampus the expression of fosB (an early gene that directs the synthesis of the FosB and DeltaFosB proteins from the AP-1 complex of transcription factors) is induced immediately after ischemia and is accompanied by an increased expression of Gal-1, especially in neurons resistant to the injury (Kurushima et al., 2005). Gal-1 induces astrocyte differentiation, with the subsequently differentiated astrocytes greatly enhancing their production of the brain-derived neurotrophic factor (BDNF) that, in turn, plays an important role in the survival, differentiation, and synaptic plasticity of neurons (Sasaki et al., 2004) (Figure 5). In contrast to the effect of oxidized Gal-1 on axonal regeneration reported above, the effects of Gal-1 on astrocyte differentiation and BDNF production depend on carbohydrate-binding activity and are astrocyte-specific since no effects on neurons have been observed (Sasaki et al., 2004). The Gal-1-triggered astrocyte differentiation occurs predominantly via a tyrosine dephosphorylation pathway that remains to be elucidated (Sasaki et al., 2004). In this context Gal-1 may thus be considered as a means for the prevention of neuronal loss in cases of injury to the central nervous system (Egnaczyk et al., 2003).

Conclusions and prospects for therapeutical applications

Gal-1 plays a number of crucial roles in (1) neuronal cell differentiation and survival in both the central and the peripheral nervous systems, and (2) the establishment and maintenance of T-cell tolerance and homeostasis in vivo. Furthermore, it is reasonable to state that Gal-1 expression or overexpression in tumors or the tissue surrounding them must be considered as a sign of their malignant progression, which is often related to the long-range dissemination of tumoral cells (metastasis), to their dissemination into the surrounding normal tissue, and to tumor immune-escape. Its increased expression is therefore associated with poor prognoses for large numbers of cancer patients.

Gal-1 could constitute a target for the setting up of novel treatments for a number of different diseases. The targeted overexpression (or delivery) of Gal-1 should be considered as a novel approach for the treatment of inflammation-related diseases including GVHD (Baum et al., 2003), arthritis (Rabinovich, Daly et al., 1999), colitis (Santucci et al., 2003), and nephritis (Tsuchiyama et al., 2000), for example. It could be also viewed as a potential therapeutic target in some neurodegenerative pathologies (Chang-Hong et al., 2005; Kadoya and Horie, 2005) and muscular dystrophies (Goldring et al., 2002). In the fight against cancer progression what should be developed for therapeutical applications is the targeted inhibition of Gal-1 expression. Indeed, the knock-down of the expression of Gal-1 in migrating tumor cells, or at least in gliomas (Lefranc et al., 2005; Camby et al., 2006) and melanomas (our unpublished results), could impair malignancy development in different ways, including, for example, a delay in cancer cell migration within the host tissue (as for gliomas in the brain) or at a distance (melanoma metastases), and the sensitization of migrating cancer cells to apoptosis. Indeed, migrating cancer cells are protected against apoptosis (Lefranc et al., 2005; Decaestecker et al., forthcoming). Restricting the migration of cancer cells by down-expressing the Gal-1 in them restores a certain level of sensitivity to cell death, and so to cytotoxic drugs (Lefranc et al., 2005; Camby et al., 2006; Decaestecker et al., forthcoming). Anti-Gal-1 compounds are thus required to combat migrating cancer cells and several groups (Andre et al., 2001; Nangia-Makker et al., 2002; Sorme et al., 2003), including our own (Ingrassia et al., 2006), are engaged in this quest. The possibility exists that such anti-Gal-1 compounds could be assayed in clinical trials (in association with cytotoxic agents) in the near future.

Conflict of interest statement

None declared.

Acknowledgments

This work enjoys the support of the “Fonds de la Recherche Scientifique Médicale” (FRSM, Belgium) and the “Fonds Yvonne Boël.” Funding to pay the Open Access publication charges for this article was provided by the “Fonds Yvonne Boël.” ML is the holder of a “Grant Télévie” from the “Fonds National de la Recherche Scientifique” (FNRS, Belgium); FL is a Clinical Research Fellow with the FNRS and RK a Director of Research with the FNRS.

References

Adams
,
L.
, Scott, G.K., and Weinberg, C.S. (
1996
) Biphasic modulation of cell growth by recombinant human galectin-1.
Biochim. Biophys. Acta
,
1312
,
137
–144.

Ahmad
,
N.
, Gabius, H.J., Sabesan, S., Oscarson, S., and Brewer, C.F. (
2004
) Thermodynamic binding studies of bivalent oligosaccharides to galectin-1, galectin-3, and the carbohydrate recognition domain of galectin-3.
Glycobiology
,
14
,
817
–825.

Akazawa
,
C.
, Nakamura, Y., Sango, K., Horie, H., and Kohsaka, S. (
2004
) Distribution of the galectin-1 mRNA in the rat nervous system: its transient upregulation in rat facial motor neurons after facial nerve axotomy.
Neuroscience
,
125
,
171
–178.

Allione
,
A.
, Wells, V., Forni, G., Mallucci, L., and Novelli, F. (
1998
) Beta-galactoside-binding protein (beta GBP) alters the cell cycle, up-regulates expression of the alpha- and beta-chains of the IFN-gamma receptor, and triggers IFN-gamma-mediated apoptosis of activated human T lymphocytes.
J. Immunol
.,
161
,
2114
–2119.

Almkvist
,
J.
, Dahlgren, C., Leffler, H., and Karlsson, A. (
2002
) Activation of the neutrophil nicotinamide adenine dinucleotide phosphate oxidase by galectin-1.
J. Immunol
.,
168
,
4034
–4041.

Almkvist
,
J.
and Karlsson, A. (
2004
) Galectins as inflammatory mediators.
Glycoconj. J
.,
19
,
575
–581.

Amano
,
M.
, Galvan, M., He, J., and Baum, L.G. (
2003
) The ST6Gal I sialyltransferase selectively modifies N-glycans on CD45 to negatively regulate galectin-1-induced CD45 clustering, phosphatase modulation, and T cell death.
J. Biol. Chem
.,
278
,
7469
–7475.

Andersen
,
H.
, Jensen, O.N., Moiseeva, E.P., and Eriksen, E.F. (
2003
) A proteome study of secreted prostatic factors affecting osteoblastic activity: galectin-1 is involved in differentiation of human bone marrow stromal cells.
J. Bone Miner. Res
.,
18
,
195
–203.

Andre
,
S.
, Kaltner, H., Furuike, T., Nishimura, S., and Gabius, H.J. (
2004
) Persubstituted cyclodextrin-based glycoclusters as inhibitors of protein-carbohydrate recognition using purified plant and mammalian lectins and wild-type and lectin-gene-transfected tumor cells as targets.
Bioconjug. Chem
.,
15
,
87
–98.

Andre
,
S.
, Pieters, R.J., Vrasidas, I., Kaltner, H., Kuwabara, I., Liu, F.T., Liskamp, R.M., and Gabius, H.J. (
2001
) Wedgelike glycodendrimers as inhibitors of binding of mammalian galectins to glycoproteins, lactose maxiclusters, and cell surface glycoconjugates.
Chembiochem
.,
2
,
822
–830.

Ashery
,
U.
, Yizhar, O., Rotblat, B., Elad-Sfadia, G., Barkan, B., Haklai, R., and Kloog, Y. (
2006
) Spatiotemporal organization of Ras signalling: rasosomes and the galectin switch.
Cell Mol. Neurobiol
. doi: 10.1007/s10571-006-9059-3.

Avni
,
O.
, Pur, Z., Yefenof, E., and Baniyash, M. (
1998
) Complement receptor, 3 of macrophages is associated with galectin-1-like protein.
J. Immunol
.,
160
,
6151
–6158.

Barondes
,
S.H.
, Castronovo, V., Cooper, D.N., Cummings, R.D., Drickamer, K., Feizi, T., Gitt, M.A., Hirabayashi, J., Hughes, C., Kasai, K., and others. (
1994
) Galectins: a family of animal beta-galactoside-binding lectins.
Cell
,
76
,
597
–598.

Bättig
,
P.
, Saudan, P., Gunde, T., and Bachmann, M.F. (
2004
) Enhanced apoptotic activity of a structurally optimized form of galectin-1.
Mol. Immunol
.,
41
,
9
–18.

Baum
,
L.G.
, Blackall, D.P., Arias-Magallano, S., Nanigian, D., Uh, S.Y., Browne, J.M., Hoffmann, D., Emmanouilides, C.E., Territo, M.C., and Baldwin, G.C. (
2003
) Amelioration of graft versus host disease by galectin-1.
Clin. Immunol
.,
109
,
295
–307.

Baum
,
L.G.
, Seilhamer, J.J., Pang, M., Levine, W.B., Beynon, D., and Berliner, J.A. (
1995
) Synthesis of an endogeneous lectin, galectin-1, by human endothelial cells is up-regulated by endothelial cell activation.
Glycoconj. J
.,
12
,
63
–68.

Berberat
,
P.O.
, Friess, H., Wang, L., Zhu, Z., Bley, T., Frigeri, L., Zimmermann, A., and Buchler, M.W. (
2001
) Comparative analysis of galectins in primary tumors and tumor metastasis in human pancreatic cancer.
J. Histochem. Cytochem
.,
49
,
539
–549.

Blaser
,
C.
, Kaufmann, M., Muller, C., Zimmermann, C., Wells, V., Mallucci, L., and Pircher, H. (
1998
) Beta-galactoside-binding protein secreted by activated T cells inhibits antigen-induced proliferation of T cells.
Eur. J. Immunol
.,
28
,
2311
–2319.

Bozic
,
M.
, Petronijevic, M., Milenkovic, S., Atanackovic, J., Lazic, J., and Vicovac, L. (
2004
) Galectin-1 and galectin-3 in the trophoblast of the gestational trophoblastic disease.
Placenta
,
25
,
797
–802.

Broker
,
L.E.
, Kruyt, F.A., and Giaccone, G. (
2005
) Cell death independent of caspases: a review.
Clin. Cancer Res
.,
11
,
3155
–3162.

Camby
,
I.
, Belot, N., Lefranc, F., Sadeghi, N., de Launoit, Y., Kaltner, H., Musette, S., Darro, F., Danguy, A., Salmon, I., and others. (
2002
) Galectin-1 modulates human glioblastoma cell migration into the brain through modifications to the actin cytoskeleton and levels of expression of small GTPases.
J. Neuropathol. Exp. Neurol
.,
61
,
585
–596.

Camby
,
I.
, Belot, N., Rorive, S., Lefranc, F., Maurage, C.A., Lahm, H., Kaltner, H., Hadari, Y., Ruchoux, M.M., Brotchi, J., and others. (
2001
) Galectins are differentially expressed in supratentorial pilocytic astrocytomas, astrocytomas, anaplastic astrocytomas and glioblastomas, and significantly modulate tumor astrocyte migration.
Brain Pathol
.,
11
,
12
–26.

Camby
,
I.
, Decaestecker, C., Lefranc, F., Kaltner, H., Gabius, H.J., and Kiss, R. (
2005
) Galectin-1 knocking down in human U87 glioblastoma cells alters their gene expression pattern.
Biochem. Biophys. Res. Commun
.,
335
,
27
–35.

Camby
,
I.
, Lefranc, F., and Kiss, R. (
2006
) In vivo siRNA-induced decreases of galectin-1 expression in experimental models of human orthotopic glioblastoma xenografts contribute therapeutic benefits. In
97th Annual Meeting of the American Association for Cancer Research (AACR),
no. 2186. Washington, DC.

Carlow
,
D.A.
, Williams, M.J., and Ziltener, H.J. (
2003
) Modulation of O‐glycans and N-glycans on murine CD8 T cells fails to alter annexin V ligand induction by galectin 1.
J. Immunol
.,
171
,
5100
–5106.

Carsons
,
S.
, Lavietes, B.B., Slomiany, A., Diamond, H.S., and Berkowitz, E. (
1987
) Carbohydrate heterogeneity of fibronectins. Synovial fluid fibronectin resembles the form secreted by cultured synoviocytes but differs from the plasma form.
J. Clin. Invest
.,
80
,
1342
–1349.

Chadli
,
A.
, LeCaer, J.P., Bladier, D., Joubert-Caron, R., and Caron, M. (
1997
) Purification and characterization of a human brain galectin-1 ligand.
J. Neurochem
.,
68
,
1640
–1647.

Chang-Hong
,
R.
, Wada, M., Koyama, S., Kimura, H., Arawaka, S., Kawanami, T., Kurita, K., Kadoya, T., Aoki, M., Itoyama, Y., and others. (
2005
) Neuroprotective effect of oxidized galectin-1 in a transgenic mouse model of amyotrophic lateral sclerosis.
Exp. Neurol
.,
194
,
203
–211.

Chiariotti
,
L.
, Benvenuto, G., Salvatore, P., Veneziani, B.M., Villone, G., Fusco, A., Russo, T., and Bruni, C.B. (
1994
) Expression of the soluble lectin L-14 gene is induced by TSH in thyroid cells and suppressed by retinoic acid in transformed neural cells.
Biochem. Biophys. Res. Commun
.,
199
,
540
–546.

Chiariotti
,
L.
, Salvatore, P., Frunzio, R., and Bruni, C.B. (
2004
) Galectin genes: regulation of expression.
Glycoconj. J
.,
19
,
441
–449.

Cho
,
M.
and Cummings, R.D. (
1995
) Galectin-1, a beta-galactoside-binding lectin in Chinese hamster ovary cells. I. Physical and chemical characterization.
J. Biol. Chem
.,
270
,
5198
–5206.

Choe
,
Y.S.
, Shim, C., Choi, D., Lee, C.S., Lee, K.K., and Kim, K. (
1997
) Expression of galectin-1 mRNA in the mouse uterus is under the control of ovarian steroids during blastocyst implantation.
Mol. Reprod. Dev
.,
48
,
261
–266.

Choufani
,
G.
, Nagy, N., Saussez, S., Marchant, H., Bisschop, P., Burchert, M., Danguy, A., Louryan, S., Salmon, I., Gabius, H.J., and others. (
1999
) The levels of expression of galectin-1, galectin-3, and the Thomsen-Friedenreich antigen and their binding sites decrease as clinical aggressiveness increases in head and neck cancers.
Cancer
,
86
,
2353
–2363.

Chung
,
C.D.
, Patel, V.P., Moran, M., Lewis, L.A., and Miceli, M.C. (
2000
) Galectin-1 induces partial TCR zeta-chain phosphorylation and antagonizes processive TCR signal transduction.
J. Immunol
.,
165
,
3722
–3729.

Cindolo
,
L.
, Benvenuto, G., Salvatore, P., Pero, R., Salvatore, G., Mirone, V., Prezioso, D., Altieri, V., Bruni, C.B., and Chiariotti, L. (
1999
) Galectin-1 and galectin-3 expression in human bladder transitional-cell carcinomas.
Int. J. Cancer
,
84
,
39
–43.

Clausse
,
N.
, van den Brule, F., Waltregny, D., Garnier, F., and Castronovo, V. (
1999
) Galectin-1 expression in prostate tumor-associated capillary endothelial cells is increased by prostate carcinoma cells and modulates heterotypic cell-cell adhesion.
Angiogenesis
,
3
,
317
–325.

Clerch
,
L.B.
, Whitney, P., Hass, M., Brew, K., Miller, T., Werner, R., and Massaro, D. (
1988
) Sequence of a full-length cDNA for rat lung beta‐galactoside-binding protein: primary and secondary structure of the lectin.
Biochemistry
,
27
,
692
–629.

Cooper
,
D.N.
(
2002
) Galectinomics: finding themes in complexity.
Biochim. Biophys. Acta
,
1572
,
209
–231.

Cooper
,
D.N.
and Barondes, S.H. (
1990
) Evidence for export of a muscle lectin from cytosol to extracellular matrix and for a novel secretory mechanism.
J. Cell. Biol
.,
110
,
1681
–1691.

Cooper
,
D.N.
, Massa, S.M., and Barondes, S.H. (
1991
) Endogenous muscle lectin inhibits myoblast adhesion to laminin.
J. Cell. Biol
.,
115
,
1437
–1448.

D’Haene
,
N.
, Maris, C., Sandras, F., Dehou, M.F., Remmelink, M., Decaestecker, C., and Salmon, I. (
2005
) The differential expression of galectin-1 and galectin-3 in normal lymphoid tissue and non-Hodgkins and Hodgkins lymphomas.
Int. J. Immunopathol. Pharmacol
.,
18
,
431
–443.

Danguy
,
A.
, Camby, I., and Kiss, R. (
2002
) Galectins and cancer.
Biochim. Biophys. Acta
,
1572
,
285
–293.

Decaestecker
,
C.
, Debeir, O., Van Ham, P., and Kiss, R. (forthcoming) Can anti-migratory drugs be screened in vitro? A review of 2D and 3D assays for the quantitative analysis of cell migration (review).
Med. Res. Rev
.

Delbrouck
,
C.
, Doyen, I., Belot, N., Decaestecker, C., Ghanooni, R., de Lavareille, A., Kaltner, H., Choufani, G., Danguy, A., Vandenhoven, G., and others. (
2002
) Galectin-1 is overexpressed in nasal polyps under budesonide and inhibits eosinophil migration.
Lab. Invest
.,
82
,
147
–158.

Dettin
,
L.
, Rubinstein, N., Aoki, A., Rabinovich, G.A., and Maldonado, C.A. (
2003
) Regulated expression and ultrastructural localization of galectin-1, a proapoptotic beta-galactoside-binding lectin, during spermatogenesis in rat testis.
Biol. Reprod
.,
68
,
51
–59.

Di Virgilio
,
S.
, Glushka, J., Moremen, K., and Pierce, M. (
1999
) Enzymatic synthesis of natural and, 13C enriched linear poly-N-acetyllactosamines as ligands for galectin-1.
Glycobiology
,
9
,
353
–364.

Dias-Baruffi
,
M.
, Zhu, H., Cho, M., Karmakar, S., McEver, R.P., and Cummings, R.D. (
2003
) Dimeric galectin-1 induces surface exposure of phosphatidylserine and phagocytic recognition of leukocytes without inducing apoptosis.
J. Biol. Chem
.,
278
,
41282
–41293.

Egnaczyk
,
G.F.
, Pomonis, J.D., Schmidt, J.A., Rogers, S.D., Peters, C., Ghilardi, J.R., Mantyh, P.W., and Maggio, J.E. (
2003
) Proteomic analysis of the reactive phenotype of astrocytes following endothelin-1 exposure.
Proteomics
,
3
,
689
–398.

Elad-Sfadia
,
G.
, Haklai, R., Ballan, E., Gabius, H.J., and Kloog, Y. (
2002
) Galectin-1 augments Ras activation and diverts Ras signals to Raf-1 at the expense of phosphoinositide 3-kinase.
J. Biol. Chem
.,
277
,
37169
–37175.

Ellerhorst
,
J.
, Nguyen, T., Cooper, D.N., Estrov, Y., Lotan, D., and Lotan, R. (
1999
) Induction of differentiation and apoptosis in the prostate cancer cell line LNCaP by sodium butyrate and galectin-1.
Int. J. Oncol
.,
14
,
225
–232.

Ellerhorst
,
J.
, Nguyen, T., Cooper, D.N., Lotan, D., and Lotan, R. (
1999
) Differential expression of endogenous galectin-1 and galectin-3 in human prostate cancer cell lines and effects of overexpressing galectin-1 on cell phenotype.
Int. J. Oncol
.,
14
,
217
–224.

Endharti
,
A.T.
, Zhou, Y.W., Nakashima, I., and Suzuki, H. (
2005
) Galectin-1 supports survival of naive T cells without promoting cell proliferation.
Eur. J. Immunol
.,
35
,
86
–97.

Fajka-Boja
,
R.
, Szemes, M., Ion, G., Legradi, A., Caron, M., and Monostori, E. (
2002
) Receptor tyrosine phosphatase, CD45 binds galectin-1 but does not mediate its apoptotic signal in T cell lines.
Immunol. Lett
.,
82
,
149
–154.

Fischer
,
C.
, Sanchez-Ruderisch, H., Welzel, M., Wiedenmann, B., Sakai, T., Andre, S., Gabius, H.J., Khachigian, L., Detjen, K.M., and Rosewicz, S. (
2005
) Galectin-1 interacts with the {alpha}5{beta}1 fibronectin receptor to restrict carcinoma cell growth via induction of p21 and p27.
J. Biol. Chem
.,
280
,
37266
–37277.

Fitzner
,
B.
, Walzel, H., Sparmann, G., Emmrich, J., Liebe, S., and Jaster, R. (
2005
) Galectin-1 is an inductor of pancreatic stellate cell activation.
Cell Signal
.,
17
,
1240
–1247.

Fouillit
,
M.
, Joubert-Caron, R., Poirier, F., Bourin, P., Monostori, E., Levi-Strauss, M., Raphael, M., Bladier, D., and Caron, M. (
2000
) Regulation of CD45-induced signaling by galectin-1 in Burkitt lymphoma B cells.
Glycobiology
,
10
,
413
–419.

Francois
,
C.
, van Velthoven, R., De Lathouwer, O., Moreno, C., Peltier, A., Kaltner, H., Salmon, I., Gabius, H.J., Danguy, A., Decaestecker, C., and others. (
1999
) Galectin-1 and galectin-3 binding pattern expression in renal cell carcinomas.
Am. J. Clin. Pathol
.,
112
,
194
–203.

Frisch
,
S.M.
and Ruoslahti, E. (
1997
) Integrins and anoikis.
Curr. Opin. Cell Biol
.,
9
,
701
–706.

Fujiwara
,
S.
, Shinkai, H., Deutzmann, R., Paulsson, M., and Timpl, R. (
1988
) Structure and distribution of N-linked oligosaccharide chains on various domains of mouse tumour laminin.
Biochem. J
.,
252
,
453
–461.

Fukaya
,
K.
, Hasegawa, M., Mashitani, T., Kadoya, T., Horie, H., Hayashi, Y., Fujisawa, H., Tachibana, O., Kida, S., and Yamashita, J. (
2003
) Oxidized galectin-1 stimulates the migration of Schwann cells from both proximal and distal stumps of transected nerves and promotes axonal regeneration after peripheral nerve injury.
J. Neuropathol. Exp. Neurol
.,
62
,
162
–172.

Gabius
,
H.J.
, Andre, S., Gunsenhauser, I., Kaltner, H., Kayser, G., Kopitz, J., Lahm, H., Harms, D., Szymas, J., and Kayser, K. (
2002
) Association of galectin-1- but not galectin-3-dependent parameters with proliferation activity in human neuroblastomas and small cell lung carcinomas.
Anticancer Res
.,
22
,
405
–410.

Galvan
,
M.
, Tsuboi, S., Fukuda, M., and Baum, L.G. (
2000
) Expression of a specific glycosyltransferase enzyme regulates T cell death mediated by galectin-1.
J. Biol. Chem
.,
275
,
16730
–16737.

Gaudier
,
E.
, Forestier, L., Gouyer, V., Huet, G., Julien, R., and Hoebler, C. (
2004
) Butyrate regulation of glycosylation-related gene expression: evidence for galectin-1 upregulation in human intestinal epithelial goblet cells.
Biochem. Biophys. Res. Commun
.,
325
,
1044
–1051.

Gauthier
,
L.
, Rossi, B., Roux, F., Termine, E., and Schiff, C. (
2002
) Galectin-1 is a stromal cell ligand of the pre-B cell receptor (BCR) implicated in synapse formation between pre-B and stromal cells and in pre-BCR triggering.
Proc. Natl. Acad. Sci. U.S.A
.,
99
,
13014
–13019.

Gillenwater
,
A.
, Xu, X.C., el-Naggar, A.K., Clayman, G.L., and Lotan, R. (
1996
) Expression of galectins in head and neck squamous cell carcinoma.
Head Neck
,
18
,
422
–432.

Gillenwater
,
A.
, Xu, X.C., Estrov, Y., Sacks, P.G., Lotan, D., and Lotan, R. (
1998
) Modulation of galectin-1 content in human head and neck squamous carcinoma cells by sodium butyrate.
Int. J. Cancer
,
75
,
217
–224.

Giordanengo
,
L.
, Gea, S., Barbieri, G., and Rabinovich, G.A. (
2001
) Anti-galectin-1 autoantibodies in human Trypanosoma cruzi infection: differential expression of this beta-galactoside-binding protein in cardiac Chagas’ disease.
Clin. Exp. Immunol
.,
124
,
266
–273.

Glinsky
,
V.V.
, Huflejt, M.E., Glinsky, G.V., Deutscher, S.L., and Quinn, T.P. (
2000
) Effects of Thomsen-Friedenreich antigen-specific peptide P-30 on beta-galactoside-mediated homotypic aggregation and adhesion to the endothelium of MDA-MB-435 human breast carcinoma cells.
Cancer Res
.,
60
,
2584
–2588.

Goldring
,
K.
, Jones, G.E., Thiagarajah, R., and Watt, D.J. (
2002
) The effect of galectin-1 on the differentiation of fibroblasts and myoblasts in vitro.
J. Cell Sci
.,
115
,
355
–366.

Grutzmann
,
R.
, Pilarsky, C., Ammerpohl, O., Luttges, J., Bohme, A., Sipos, B., Foerder, M., Alldinger, I., Jahnke, B., Schackert, H.K., and others. (
2004
) Gene expression profiling of microdissected pancreatic ductal carcinomas using high-density DNA microarrays.
Neoplasia
,
6
,
611
–622.

Gu
,
J.
and Taniguchi, N. (
2004
) Regulation of integrin functions by N-glycans.
Glycoconj. J
.,
21
,
9
–15.

Gu
,
M.
, Wang, W., Song, W.K., Cooper, D.N., and Kaufman, S.J. (
1994
) Selective modulation of the interaction of alpha, 7 beta 1 integrin with fibronectin and laminin by L-14 lectin during skeletal muscle differentiation.
J. Cell Sci
.,
107
,
175
–181.

Gunnersen
,
J.M.
, Spirkoska, V., Smith, P.E., Danks, R.A., and Tan, S.S. (
2000
) Growth and migration markers of rat C6 glioma cells identified by serial analysis of gene expression.
Glia
,
32
,
146
–154.

Hahn
,
H.P.
, Pang, M., He, J., Hernandez, J.D., Yang, R.Y., Li, L.Y., Wang, X., Liu, F.T., and Baum, L.G. (
2004
) Galectin-1 induces nuclear translocation of endonuclease G in caspase- and cytochrome c‐independent T cell death.
Cell Death Differ
.,
11
,
1277
–1286.

Hansen
,
T.V.
, Hammer, N.A., Nielsen, J., Madsen, M., Dalbaeck, C., Wewer, U.M., Christiansen, J., and Nielsen, F.C. (
2004
) Dwarfism and impaired gut development in insulin-like growth factor II mRNA-binding protein 1-deficient mice.
Mol. Cell Biol
.,
24
,
4448
–4464.

Harrison
,
F.L.
and Wilson, T.J. (
1992
) The 14 kDa beta-galactoside binding lectin in myoblast and myotube cultures: localization by confocal microscopy.
J. Cell Sci
.,
101
,
635
–646.

Harvey
,
S.
, Zhang, Y., Landry, F., Miller, C., and Smith, J.W. (
2001
) Insights into a plasma membrane signature.
Physiol. Genomics
,
5
,
129
–136.

He
,
J.
and Baum, L.G. (
2004
) Presentation of galectin-1 by extracellular matrix triggers T cell death.
J. Biol. Chem
.,
279
,
4705
–4712.

He
,
Q.Y.
, Chen, J., Kung, H.F., Yuen, A.P., and Chiu, J.F. (
2004
) Identification of tumor-associated proteins in oral tongue squamous cell carcinoma by proteomics.
Proteomics
,
4
,
271
–278.

Hirabayashi
,
J.
, Ayaki, H., Soma, G., and Kasai, K. (
1989
) Cloning and nucleotide sequence of a full-length cDNA for human 14 kDa beta-galactoside-binding lectin.
Biochim. Biophys. Acta
,
1008
,
85
–91.

Hittelet
,
A.
, Legendre, H., Nagy, N., Bronckart, Y., Pector, J.C., Salmon, I., Yeaton, P., Gabius, H.J., Kiss, R., and Camby, I. (
2003
) Upregulation of galectins-1 and -3 in human colon cancer and their role in regulating cell migration.
Int. J. Cancer
103
,
370
–379.

Horie
,
H.
, Inagaki, Y., Sohma, Y., Nozawa, R., Okawa, K., Hasegawa, M., Muramatsu, N., Kawano, H., Horie, M., Koyama, H., and others. (
1999
) Galectin-1 regulates initial axonal growth in peripheral nerves after axotomy.
J. Neurosci
.,
19
,
9964
–9974.

Horie
,
H.
and Kadoya, T. (
2000
) Identification of oxidized galectin-1 as an initial repair regulatory factor after axotomy in peripheral nerves.
Neurosci. Res
.,
38
,
131
–137.

Horie
,
H.
, Kadoya, T., Hikawa, N., Sango, K., Inoue, H., Takeshita, K., Asawa, R., Hiroi, T., Sato, M., Yoshioka, T., and others. (
2004
) Oxidized galectin-1 stimulates macrophages to promote axonal regeneration in peripheral nerves after axotomy.
J. Neurosci
.,
24
,
1873
–1880.

Horiguchi
,
N.
, Arimoto, K., Mizutani, A., Endo-Ichikawa, Y., Nakada, H., and Taketani, S. (
2003
) Galectin-1 induces cell adhesion to the extracellular matrix and apoptosis of non-adherent human colon cancer Colo201 cells.
J. Biochem. (Tokyo)
,
134
,
869
–874.

Hughes
,
R.C.
(
1999
) Secretion of the galectin family of mammalian carbohydrate-binding proteins.
Biochim. Biophys. Acta
,
1473
,
172
–185.

Hughes
,
R.C.
(
2004
) Galectins in kidney development.
Glycoconj. J
.,
19
,
621
–629.

Hutvagner
,
G.
and Zamore, P.D. (
2002
) A microRNA in a multiple-turnover RNAi enzyme complex.
Science
,
297
,
2056
–2060.

Hynes
,
M.A.
, Gitt, M., Barondes, S.H., Jessell, T.M., and Buck, L.B. (
1990
) Selective expression of an endogenous lactose-binding lectin gene in subsets of central and peripheral neurons.
J. Neurosci
.,
10
,
1004
–1013.

Inagaki
,
Y.
, Sohma, Y., Horie, H., Nozawa, R., and Kadoya, T. (
2000
) Oxidized galectin-1 promotes axonal regeneration in peripheral nerves but does not possess lectin properties.
Eur. J. Biochem
.,
267
,
2955
–2964.

Ingrassia
,
L.
, Nshimyumukiza, P., Dewelle, J., Lefranc, F., Wlodarczak, L., Thomas, S., Dielie, G., Chiron, C., Zedde, C., Tisnès, P., and others. (
2006
) A lactosylated steroid contributes in vivo therapeutic benefits in experimental models of mouse lymphoma and human glioblastoma.
J. Med. Chem
.,
49
,
1800
–1807.

Ion
,
G.
, Fajka-Boja, R., Toth, G.K., Caron, M., and Monostori, E. (
2005
) Role of p56lck and ZAP70-mediated tyrosine phosphorylation in galectin-1-induced cell death.
Cell Death Differ
.,
12
,
1145
–1147.

Jeschke
,
U.
, Reimer, T., Bergemann, C., Wiest, I., Schulze, S., Friese, K., and Walzel, H. (
2004
) Binding of galectin-1 (gal-1) on trophoblast cells and inhibition of hormone production of trophoblast tumor cells in vitro by gal-1.
Histochem. Cell Biol
.,
121
,
501
–508.

Joubert
,
R.
, Caron, M., Avellana-Adalid, V., Mornet, D., and Bladier, D. (
1992
) Human brain lectin: a soluble lectin that binds actin.
J. Neurochem
.,
58
,
200
–203.

Joubert
,
R.
, Kuchler, S., Zanetta, J.P., Bladier, D., Avellana-Adalid, V., Caron, M., Doinel, C., and Vincendon, G. (
1989
) Immunohistochemical localization of a beta-galactoside-binding lectin in rat central nervous system. I. Light- and electron-microscopical studies on developing cerebral cortex and corpus callosum.
Dev. Neurosci
.,
11
,
397
–413.

Kadoya
,
T.
and Horie, H. (
2005
) Structural and functional studies of galectin-1: a novel axonal regeneration-promoting activity for oxidized galectin-1.
Curr. Drug Targets
,
6
,
375
–383.

Kadoya
,
T.
, Oyanagi, K., Kawakami, E., Hasegawa, M., Inagaki, Y., Sohma, Y., and Horie, H. (
2005
) Oxidized galectin-1 advances the functional recovery after peripheral nerve injury.
Neurosci. Lett
.,
380
,
284
–288.

Kondoh
,
N.
, Hada, A., Ryo, A., Shuda, M., Arai, M., Matsubara, O., Kimura, F., Wakatsuki, T., and Yamamoto, M. (
2003
) Activation of galectin-1 gene in human hepatocellular carcinoma involves methylation-sensitive complex formations at the transcriptional upstream and downstream elements.
Int. J. Oncol
.,
23
,
1575
–1583.

Kopitz
,
J.
, von Reitzenstein, C., Andre, S., Kaltner, H., Uhl, J., Ehemann, V., Cantz, M., and Gabius, H.J. (
2001
) Negative regulation of neuroblastoma cell growth by carbohydrate-dependent surface binding of galectin-1 and functional divergence from galectin-3.
J. Biol. Chem
.,
276
,
35917
–35923.

Kopitz
,
J.
, von Reitzenstein, C., Burchert, M., Cantz, M., and Gabius, H.J. (
1998
) Galectin-1 is a major receptor for ganglioside GM1, a product of the growth-controlling activity of a cell surface ganglioside sialidase, on human neuroblastoma cells in culture.
J. Biol. Chem
.,
273
,
11205
–11211.

Kultima
,
K.
, Nystrom, A.M., Scholz, B., Gustafson, A.L., Dencker, L., and Stigson, M. (
2004
) Valproic acid teratogenicity: a toxicogenomics approach.
Environ. Health Perspect
.,
112
,
1225
–12235.

Kurushima
,
H.
, Ohno, M., Miura, T., Nakamura, T.Y., Horie, H., Kadoya, T., Ooboshi, H., Kitazono, T., Ibayashi, S., Iida, M., and others. (
2005
) Selective induction of DeltaFosB in the brain after transient forebrain ischemia accompanied by an increased expression of galectin-1, and the implication of DeltaFosB and galectin-1 in neuroprotection and neurogenesis.
Cell Death Differ
.,
12
,
1078
–1096.

La
,
M.
, Cao, T.V., Cerchiaro, G., Chilton, K., Hirabayashi, J., Kasai, K., Oliani, S.M., Chernajovsky, Y., and Perretti, M. (
2003
) A novel biological activity for galectin-1: inhibition of leukocyte–endothelial cell interactions in experimental inflammation.
Am. J. Pathol
.,
163
,
1505
–1515.

Lanteri
,
M.
, Giordanengo, V., Hiraoka, N., Fuzibet, J.G., Auberger, P., Fukuda, M., Baum, L.G., and Lefebvre, J.C. (
2003
) Altered T cell surface glycosylation in HIV-1 infection results in increased susceptibility to galectin-1-induced cell death.
Glycobiology
,
13
,
909
–918.

Le
,
Q.T.
, Shi, G., Cao, H., Nelson, D.W., Wang, Y., Chen, E.Y., Zhao, S., Kong, C., Richardson, D., O’Byrne, K., J., and others. (
2005
) Galectin-1: a link between tumor hypoxia and tumor immune privilege.
J. Clin. Oncol
.,
23
,
8932
–8941.

Lefranc
,
F.
, Brotchi, J., and Kiss, R. (
2005
) Possible future issues in the treatment of glioblastomas: special emphasis on cell migration and the resistance of migrating glioblastoma cells to apoptosis.
J. Clin. Oncol
.,
23
,
2411
–2422.

Leppanen
,
A.
, Stowell, S., Blixt, O., and Cummings, R.D. (
2005
) Dimeric galectin-1 binds with high affinity to alpha2,3-sialylated and non-sialylated terminal N-acetyllactosamine units on surface-bound extended glycans.
J. Biol. Chem
.,
280
,
5549
–5562.

Levi
,
G.
, Tarrab-Hazdai, R., and Teichberg, V.I. (
1983
) Prevention and therapy with electrolectin of experimental autoimmune myasthenia gravis in rabbits.
Eur. J. Immunol
.,
13
,
500
–507.

Levroney
,
E.L.
, Aguilar, H.C., Fulcher, J.A., Kohatsu, L., Pace, K.E., Pang, M., Gurney, K.B., Baum, L.G., and Lee, B. (
2005
) Novel innate immune functions for galectin-1: galectin-1 inhibits cell fusion by Nipah virus envelope glycoproteins and augments dendritic cell secretion of proinflammatory cytokines.
J. Immunol
.,
175
,
413
–420.

Lim
,
J.W.
, Kim, H., and Kim, K.H. (
2003
) Cell adhesion-related gene expression by Helicobacter pylori in gastric epithelial AGS cells.
Int. J. Biochem. Cell. Biol
.,
35
,
1284
–1296.

Liu
,
F.T.
(
2000
) Galectins: a new family of regulators of inflammation.
Clin. Immunol
.,
97
,
79
–88.

Liu
,
F.T.
and Rabinovich, G.A. (
2005
) Galectins as modulators of tumour progression.
Nat. Rev. Cancer
,
5
,
29
–41.

Lopez-Lucendo
,
M.F.
, Solis, D., Andre, S., Hirabayashi, J., Kasai, K., Kaltner, H., Gabius, H.J., and Romero, A. (
2004
) Growth-regulatory human galectin-1: crystallographic characterisation of the structural changes induced by single-site mutations and their impact on the thermodynamics of ligand binding.
J. Mol. Biol
.,
343
,
957
–970.

Lu
,
Y.
, Lotan, D., and Lotan, R. (
2000
) Differential regulation of constitutive and retinoic acid-induced galectin-1 gene transcription in murine embryonal carcinoma and myoblastic cells.
Biochim. Biophys. Acta
,
1491
,
13
–19.

Lu
,
Y.
and Lotan, R. (
1999
) Transcriptional regulation by butyrate of mouse galectin-1 gene in embryonal carcinoma cells.
Biochim. Biophys. Acta
,
1444
,
85
–91.

Lutomski
,
D.
, Fouillit, M., Bourin, P., Mellottee, D., Denize, N., Pontet, M., Bladier, D., Caron, M., and Joubert-Caron, R. (
1997
) Externalization and binding of galectin-1 on cell surface of K562 cells upon erythroid differentiation.
Glycobiology
,
7
,
1193
–1199.

Lutomski
,
D.
, Joubert-Caron, R., Lefebure, C., Salama, J., Belin, C., Bladier, D., and Caron, M. (
1997
) Anti-galectin-1 autoantibodies in serum of patients with neurological diseases.
Clin. Chim. Acta
,
262
,
131
–138.

Mackay
,
A.
, Jones, C., Dexter, T., Silva, R.L., Bulmer, K., Jones, A., Simpson, P., Harris, R.A., Jat, P.S., Neville, A.M., and others. (
2003
) cDNA microarray analysis of genes associated with ERBB2 (HER2/neu) overexpression in human mammary luminal epithelial cells.
Oncogene
,
22
,
2680
–2688.

Maeda
,
N.
, Kawada, N., Seki, S., Arakawa, T., Ikeda, K., Iwao, H., Okuyama, H., Hirabayashi, J., Kasai, K., and Yoshizato, K. (
2003
) Stimulation of proliferation of rat hepatic stellate cells by galectin-1 and galectin-3 through different intracellular signaling pathways.
J. Biol. Chem
.,
278
,
18938
–18944.

Mahanthappa
,
N.K.
, Cooper, D.N., Barondes, S.H., and Schwarting, G.A. (
1994
) Rat olfactory neurons can utilize the endogenous lectin, L‐14, in a novel adhesion mechanism.
Development
,
120
,
1373
–1384.

Maquoi
,
E.
, van den Brule, F.A., Castronovo, V., and Foidart, J.M. (
1997
) Changes in the distribution pattern of galectin-1 and galectin-3 in human placenta correlates with the differentiation pathways of trophoblasts.
Placenta
,
18
,
433
–439.

Martinez
,
V.G.
, Pellizzari, E.H., Diaz, E.S., Cigorraga, S.B., Lustig, L., Denduchis, B., Wolfenstein-Todel, C., and Iglesias, M.M. (
2004
) Galectin-1, a cell adhesion modulator, induces apoptosis of rat Leydig cells in vitro.
Glycobiology
,
14
,
127
–137.

Matarrese
,
P.
, Tinari, A., Mormone, E., Bianco, G.A., Toscano, M.A., Ascione, B., Rabinovich, G.A., and Malorni, W. (
2005
) Galectin-1 sensitizes resting human T lymphocytes to Fas (CD95) -mediated cell death via mitochondrial hyperpolarization, budding, and fission.
J. Biol. Chem
.,
280
,
6969
–6985.

McGraw
,
J.
, Gaudet, A.D., Oschipok, L.W., Kadoya, T., Horie, H., Steeves, J.D., Tetzlaff, W., and Ramer, M.S. (
2005
) Regulation of neuronal and glial galectin-1 expression by peripheral and central axotomy of rat primary afferent neurons.
Exp. Neurol
.,
195
,
103
–114.

McGraw
,
J.
, Gaudet, A.D., Oschipok, L.W., Steeves, J.D., Poirier, F., Tetzlaff, W., and Ramer, M.S. (
2005
) Altered primary afferent anatomy and reduced thermal sensitivity in mice lacking galectin-1.
Pain
,
114
,
7
–18.

McGraw
,
J.
, McPhail, L.T., Oschipok, L.W., Horie, H., Poirier, F., Steeves, J.D., Ramer, M.S., and Tetzlaff, W. (
2004
) Galectin-1 in regenerating motoneurons.
Eur. J. Neurosci
.,
20
,
2872
–2880.

McGraw
,
J.
, Oschipok, L.W., Liu, J., Hiebert, G.W., Mak, C.F., Horie, H., Kadoya, T., Steeves, J.D., Ramer, M.S., and Tetzlaff, W. (
2004
) Galectin-1 expression correlates with the regenerative potential of rubrospinal and spinal motoneurons.
Neuroscience
,
128
,
713
–719.

Miura
,
T.
, Takahashi, M., Horie, H., Kurushima, H., Tsuchimoto, D., Sakumi, K., and Nakabeppu, Y. (
2004
) Galectin-1beta, a natural monomeric form of galectin-1 lacking its six amino-terminal residues promotes axonal regeneration but not cell death.
Cell Death Differ
.,
11
,
1076
–1083.

Moiseeva
,
E.P.
, Javed, Q., Spring, E.L., and de Bono, D.P. (
2000
) Galectin 1 is involved in vascular smooth muscle cell proliferation.
Cardiovasc. Res
.,
45
,
493
–502.

Moiseeva
,
E.P.
, Spring, E.L., Baron, J.H., and de Bono, D.P. (
1999
) Galectin 1 modulates attachment, spreading and migration of cultured vascular smooth muscle cells via interactions with cellular receptors and components of extracellular matrix.
J. Vasc. Res
.,
36
,
47
–58.

Moiseeva
,
E.P.
, Williams, B., Goodall, A.H., and Samani, N.J. (
2003
) Galectin-1 interacts with beta-1 subunit of integrin.
Biochem. Biophys. Res. Commun
.,
310
,
1010
–1016.

Moiseeva
,
E.P.
, Williams, B., and Samani, N.J. (
2003
) Galectin 1 inhibits incorporation of vitronectin and chondroitin sulfate B into the extracellular matrix of human vascular smooth muscle cells.
Biochim. Biophys. Acta
,
1619
,
125
–132.

Nagy
,
N.
, Legendre, H., Engels, O., Andre, S., Kaltner, H., Wasano, K., Zick, Y., Pector, J.C., Decaestecker, C., Gabius, H.J., and others. (
2003
) Refined prognostic evaluation in colon carcinoma using immunohistochemical galectin fingerprinting.
Cancer
,
97
,
1849
–1858.

Nangia-Makker
,
P.
, Conklin, J., Hogan, V., and Raz, A. (
2002
) Carbohydrate-binding proteins in cancer, and their ligands as therapeutic agents.
Trends Mol. Med
.,
8
,
187
–192.

Nguyen
,
J.T.
, Evans, D.P., Galvan, M., Pace, K.E., Leitenberg, D., Bui, T.N., and Baum, L.G. (
2001
) CD45 modulates galectin-1-induced T cell death: regulation by expression of core 2 O-glycans.
J. Immunol
.,
167
,
5697
–5707.

Nickel
,
W.
(
2005
) Unconventional secretory routes: direct protein export across the plasma membrane of mammalian cells.
Traffic
,
6
,
607
–614.

Offner
,
H.
, Celnik, B., Bringman, T.S., Casentini-Borocz, D., Nedwin, G.E., and Vandenbark, A.A. (
1990
) Recombinant human beta-galactoside binding lectin suppresses clinical and histological signs of experimental autoimmune encephalomyelitis.
J. Neuroimmunol
.,
28
,
177
–184.

Ohannesian
,
D.W.
, Lotan, D., and Lotan, R. (
1994
) Concomitant increases in galectin-1 and its glycoconjugate ligands (carcinoembryonic antigen, lamp-1, and lamp-2) in cultured human colon carcinoma cells by sodium butyrate.
Cancer Res
.,
54
,
5992
–6000.

Ouellet
,
M.
, Mercier, S., Pelletier, I., Bounou, S., Roy, J., Hirabayashi, J., Sato, S., and Tremblay, M.J. (
2005
) Galectin-1 acts as a soluble host factor that promotes HIV-1 infectivity through stabilization of virus attachment to host cells.
J. Immunol
.,
174
,
4120
–4126.

Outenreath
,
R.L.
and Jones, A.L. (
1992
) Influence of an endogenous lectin substrate on cultured dorsal root ganglion cells.
J. Neurocytol
.,
21
,
788
–795.

Ozeki
,
Y.
, Matsui, T., Yamamoto, Y., Funahashi, M., Hamako, J., and Titani, K. (
1995
) Tissue fibronectin is an endogenous ligand for galectin-1.
Glycobiology
,
5
,
255
–261.

Pace
,
K.E.
, Hahn, H.P., Pang, M., Nguyen, J.T., and Baum, L.G. (
2000
) CD7 delivers a pro-apoptotic signal during galectin-1-induced T cell death.
J. Immunol
.,
165
,
2331
–2334.

Pace
,
K.E.
, Lee, C., Stewart, P.L., and Baum, L.G. (
1999
) Restricted receptor segregation into membrane microdomains occurs on human T cells during apoptosis induced by galectin-1.
J. Immunol
.,
163
,
3801
–3811.

Park
,
J.W.
, Voss, P.G., Grabski, S., Wang, J.L., and Patterson, R.J. (
2001
) Association of galectin-1 and galectin-3 with Gemin4 in complexes containing the SMN protein.
Nucleic Acids Res
.,
29
,
3595
–3602.

Paushkin
,
S.
, Gubitz, A.K., Massenet, S., and Dreyfuss, G. (
2002
) The SMN complex, an assemblyosome of ribonucleoproteins.
Curr. Opin. Cell Biol
.,
14
,
305
–312.

Paz
,
A.
, Haklai, R., Elad-Sfadia, G., Ballan, E., and Kloog, Y. (
2001
) Galectin-1 binds oncogenic H-Ras to mediate Ras membrane anchorage and cell transformation.
Oncogene
,
20
,
7486
–7493.

Perillo
,
N.L.
, Pace, K.E., Seilhamer, J.J., and Baum, L.G. (
1995
) Apoptosis of T cells mediated by galectin-1.
Nature
,
378
,
736
–739.

Perillo
,
N.L.
, Uittenbogaart, C.H., Nguyen, J.T., and Baum, L.G. (
1997
) Galectin-1, an endogenous lectin produced by thymic epithelial cells, induces apoptosis of human thymocytes.
J. Exp. Med
.,
185
,
1851
–1858.

Poirier
,
F.
, Bourin, P., Bladier, D., Joubert-Caron, R., and Caron, M. (
2001
) Effect of 5-azacytidine and galectin-1 on growth and differentiation of the human b lymphoma cell line bl36.
Cancer Cell. Int
.,
1
,
2
.

Poirier
,
F.
and Robertson, E.J. (
1993
) Normal development of mice carrying a null mutation in the gene encoding the L14 S-type lectin.
Development
119
,
1229
–1236.

Prior
,
I.A.
, Muncke, C., Parton, R.G., and Hancock, J.F. (
2003
) Direct visualization of Ras proteins in spatially distinct cell surface microdomains.
J. Cell Biol
.,
160
,
165
–170.

Puche
,
A.C.
, Poirier, F., Hair, M., Bartlett, P.F., and Key, B. (
1996
) Role of galectin-1 in the developing mouse olfactory system.
Dev. Biol
.,
179
,
274
–287.

Rabinovich
,
G.A.
, Alonso, C.R., Sotomayor, C.E., Durand, S., Bocco, J.L., and Riera, C.M. (
2000
) Molecular mechanisms implicated in galectin-1-induced apoptosis: activation of the AP-1 transcription factor and downregulation of Bcl-2.
Cell Death Differ
.,
7
,
747
–753.

Rabinovich
,
G.A.
, Ariel, A., Hershkoviz, R., Hirabayashi, J., Kasai, K.I., and Lider, O. (
1999
) Specific inhibition of T-cell adhesion to extracellular matrix and proinflammatory cytokine secretion by human recombinant galectin-1.
Immunology
,
97
,
100
–106.

Rabinovich
,
G.A.
, Daly, G., Dreja, H., Tailor, H., Riera, C.M., Hirabayashi, J., and Chernajovsky, Y. (
1999
) Recombinant galectin-1 and its genetic delivery suppress collagen-induced arthritis via T cell apoptosis.
J. Exp. Med
.,
190
,
385
–398.

Rabinovich
,
G.A.
, Iglesias, M.M., Modesti, N.M., Castagna, L.F., Wolfenstein-Todel, C., Riera, C.M., and Sotomayor, C.E. (
1998
) Activated rat macrophages produce a galectin-1-like protein that induces apoptosis of T cells: biochemical and functional characterization.
J. Immunol
.,
160
,
4831
–4840.

Rabinovich
,
G.A.
, Ramhorst, R.E., Rubinstein, N., Corigliano, A., Daroqui, M.C., Kier-Joffe, E.B., and Fainboim, L. (
2002
) Induction of allogenic T-cell hyporesponsiveness by galectin-1-mediated apoptotic and non-apoptotic mechanisms.
Cell Death Differ
.,
9
,
661
–670.

Rabinovich
,
G.A.
, Rubinstein, N., Matar, P., Rozados, V., Gervasoni, S., and Scharovsky, G.O. (
2002
) The antimetastatic effect of a single low dose of cyclophosphamide involves modulation of galectin-1 and Bcl-2 expression.
Cancer Immunol. Immunother
.,
50
,
597
–603.

Rabinovich
,
G.A.
, Sotomayor, C.E., Riera, C.M., Bianco, I., and Correa, S.G. (
2000
) Evidence of a role for galectin-1 in acute inflammation.
Eur. J. Immunol
.,
30
,
1331
–1339.

Rappl
,
G.
, Abken, H., Muche, J.M., Sterry, W., Tilgen, W., Andre, S., Kaltner, H., Ugurel, S., Gabius, H.J., and Reinhold, U. (
2002
) CD4+CD7- leukemic T cells from patients with Sezary syndrome are protected from galectin-1-triggered T cell death.
Leukemia
,
16
,
840
–845.

Regan
,
L.J.
, Dodd, J., Barondes, S.H., and Jessell, T.M. (
1986
) Selective expression of endogenous lactose-binding lectins and lactoseries glycoconjugates in subsets of rat sensory neurons.
Proc. Natl. Acad. Sci. U.S.A
.,
83
,
2248
–2252.

Roberts
,
A.A.
, Amano, M., Felten, C., Galvan, M., Sulur, G., Pinter-Brown, L., Dobbeling, U., Burg, G., Said, J., and Baum, L.G. (
2003
) Galectin-1-mediated apoptosis in mycosis fungoides: the roles of CD7 and cell surface glycosylation.
Mod. Pathol
.,
16
,
543
–551.

Rochard
,
P.
, Galiegue, S., Tinel, N., Peleraux, A., Bord, A., Jbilo, O., and Casellas, P. (
2004
) Expression of the peripheral benzodiazepine receptor triggers thymocyte differentiation.
Gene Expr
.,
12
,
13
–27.

Rorive
,
S.
, Belot, N., Decaestecker, C., Lefranc, F., Gordower, L., Micik, S., Maurage, C.A., Kaltner, H., Ruchoux, M.M., Danguy, A., and others. (
2001
) Galectin-1 is highly expressed in human gliomas with relevance for modulation of invasion of tumor astrocytes into the brain parenchyma.
Glia
,
33
,
241
–255.

Rotblat
,
B.
, Niv, H., Andre, S., Kaltner, H., Gabius, H.J., and Kloog, Y. (
2004
) Galectin-1 (L11A) predicted from a computed galectin-1 farnesyl-binding pocket selectively inhibits Ras-GTP.
Cancer Res
.,
64
,
3112
–3118.

Rubinstein
,
N.
, Alvarez, M., Zwirner, N.W., Toscano, M.A., Ilarregui, J.M., Bravo, A., Mordoh, J., Fainboim, L., Podhajcer, O.L., and Rabinovich, G.A. (
2004
) Targeted inhibition of galectin-1 gene expression in tumor cells results in heightened T cell-mediated rejection; a potential mechanism of tumor-immune privilege.
Cancer Cell
,
5
,
241
–251.

Rubinstein
,
N.
, Ilarregui, J.M., Toscano, M.A., and Rabinovich, G.A. (
2004
) The role of galectins in the initiation, amplification and resolution of the inflammatory response.
Tissue Antigens
,
64
,
1
–12.

Sanford
,
G.L.
and Harris-Hooker, S. (
1990
) Stimulation of vascular cell proliferation by beta-galactoside specific lectins.
Faseb J
.,
4
,
2912
–2918.

Sango
,
K.
, Tokashiki, A., Ajiki, K., Horie, M., Kawano, H., Watabe, K., Horie, H., and Kadoya, T. (
2004
) Synthesis, localization and externalization of galectin-1 in mature dorsal root ganglion neurons and Schwann cells.
Eur. J. Neurosci
.,
19
,
55
–64.

Sanjuan
,
X.
, Fernandez, P.L., Castells, A., Castronovo, V., van den Brule, F., Liu, F.T., Cardesa, A., and Campo, E. (
1997
) Differential expression of galectin 3 and galectin 1 in colorectal cancer progression.
Gastroenterology
113
,
1906
–1915.

Santucci
,
L.
, Fiorucci, S., Cammilleri, F., Servillo, G., Federici, B., and Morelli, A. (
2000
) Galectin-1 exerts immunomodulatory and protective effects on concanavalin A-induced hepatitis in mice.
Hepatology
,
31
,
399
–406.

Santucci
,
L.
, Fiorucci, S., Rubinstein, N., Mencarelli, A., Palazzetti, B., Federici, B., Rabinovich, G.A., and Morelli, A. (
2003
) Galectin-1 suppresses experimental colitis in mice.
Gastroenterology
,
124
,
1381
–1394.

Sasaki
,
T.
, Hirabayashi, J., Manya, H., Kasai, K., and Endo, T. (
2004
) Galectin-1 induces astrocyte differentiation, which leads to production of brain-derived neurotrophic factor.
Glycobiology
,
14
,
357
–363.

Saussez
,
S.
, Cucu, D.R., Deacestecker, C., Chevalier, D., Kaltner, H., Andre, S., Wacreniez, A., Toubeau, G., Camby, I., Gabius, H.J., and Kiss, R. (
2006
) Galectin-7 (p53-induced gene-1): a new prognostic predictor of recurrence and survival in stage IV hypopharyngeal cancer.
Ann. Surg. Oncol
.,
13
,
999
–1009.

Saussez
,
S.
, Nonclercq, D., Laurent, G., Wattiez, R., Andre, S., Kaltner, H., Gabius, H.J., Kiss, R., and Toubeau, G. (
2005
) Toward functional glycomics by localization of tissue lectins: immunohistochemical galectin fingerprinting during diethylstilbestrol-induced kidney tumorigenesis in male Syrian hamster.
Histochem. Cell. Biol
.,
123
,
29
–41.

Savin
,
S.B.
, Cvejic, D.S., and Jankovic, M.M. (
2003
) Expression of galectin-1 and galectin-3 in human fetal thyroid gland.
J. Histochem. Cytochem
.,
51
,
479
–483.

Schwarz
,
F.P.
, Ahmed, H., Bianchet, M.A., Amzel, L.M., and Vasta, G.R. (
1998
) Thermodynamics of bovine spleen galectin-1 binding to disaccharides: correlation with structure and its effect on oligomerization at the denaturation temperature.
Biochemistry
,
37
,
5867
–5877.

Schwarz
,
G.
, Jr., Remmelink, M., Decaestecker, C., Gielen, I., Budel, V., Burchert, M., Darro, F., Danguy, A., Gabius, H.J., Salmon, I., and others. (
1999
) Galectin fingerprinting in tumor diagnosis. Differential expression of galectin-3 and galectin-3 binding sites, but not galectin-1, in benign vs malignant uterine smooth muscle tumors.
Am. J. Clin. Pathol
.,
111
,
623
–631.

Seelenmeyer
,
C.
, Wegehingel, S., Lechner, J., and Nickel, W. (
2003
) The cancer antigen CA125 represents a novel counter receptor for galectin-1.
J. Cell Sci
.,
116
,
1305
–1318.

Shen
,
J.
, Person, M.D., Zhu, J., Abbruzzese, J.L., and Li, D. (
2004
) Protein expression profiles in pancreatic adenocarcinoma compared with normal pancreatic tissue and tissue affected by pancreatitis as detected by two-dimensional gel electrophoresis and mass spectrometry.
Cancer Res
.,
64
,
9018
–9026.

Shimonishi
,
T.
, Miyazaki, K., Kono, N., Sabit, H., Tuneyama, K., Harada, K., Hirabayashi, J., Kasai, K., and Nakanuma, Y. (
2001
) Expression of endogenous galectin-1 and galectin-3 in intrahepatic cholangiocarcinoma.
Hum. Pathol
.,
32
,
302
–310.

Silva
,
W.A.
, Jr., Covas, D.T., Panepucci, R.A., Proto-Siqueira, R., Siufi, J.L., Zanette, D.L., Santos, A.R., and Zago, M.A. (
2003
) The profile of gene expression of human marrow mesenchymal stem cells.
Stem Cells
,
21
,
661
–669.

Simon
,
P.
, Decaestecker, C., Choufani, G., Delbrouck, C., Danguy, A., Salmon, I., Zick, Y., Kaltner, H., Hassid, S., Gabius, H.J., and others. (
2001
) The levels of retinoid RARbeta receptors correlate with galectin-1–3 and -8 expression in human cholesteatomas.
Hear. Res
.,
156
,
1
–9.

Sorme
,
P.
, Kahl-Knutsson, B., Wellmar, U., Magnusson, B.G., Leffler, H., and Nilsson, U.J. (
2003
) Design and synthesis of galectin inhibitors.
Methods Enzymol
.,
363
,
157
–169.

Stowell
,
S.R.
, Dias-Baruffi, M., Penttila, L., Renkonen, O., Nyame, A.K., and Cummings, R.D. (
2004
) Human galectin-1 recognition of poly-N‐acetyllactosamine and chimeric polysaccharides.
Glycobiology
,
14
,
157
–167.

Stupack
,
D.G.
and Cheresch, D.A. (
2002
) Get a ligand, get a life: integrins, signaling and cell survival.
J. Cell Sci
.,
115
,
3729
–3738.

Symons
,
A.
, Cooper, D.N., and Barclay, A.N. (
2000
) Characterization of the interaction between galectin-1 and lymphocyte glycoproteins CD45 and Thy-1.
Glycobiology
,
10
,
559
–563.

Szoke
,
T.
, Kayser, K., Baumhakel, J.D., Trojan, I., Furak, J., Tiszlavicz, L., Horvath, A., Szluha, K., Gabius, H.J., and Andre, S. (
2005
) Prognostic significance of endogenous adhesion/growth-regulatory lectins in lung cancer.
Oncology
,
69
,
167
–174.

Tahara
,
K.
, Tsuchimoto, D., Tominaga, Y., Asoh, S., Ohta, S., Kitagawa, M., Horie, H., Kadoya, T., and Nakabeppu, Y. (
2003
) DeltaFosB, but not FosB, induces delayed apoptosis independent of cell proliferation in the Rat1a embryo cell line.
Cell Death Differ
.,
10
,
496
–507.

Tenne-Brown
,
J.
, Puche, A.C., and Key, B. (
1998
) Expression of galectin-1 in the mouse olfactory system.
Int. J. Dev. Biol
.,
42
,
791
–799.

Timmons
,
P.M.
, Rigby, P.W., and Poirier, F. (
2002
) The murine seminiferous epithelial cycle is pre-figured in the Sertoli cells of the embryonic testis.
Development
,
129
,
635
–647.

Tinari
,
N.
, Kuwabara, I., Huflejt, M.E., Shen, P.F., Iacobelli, S., and Liu, F.T. (
2001
) Glycoprotein, 90K/MAC-2BP interacts with galectin-1 and mediates galectin-1-induced cell aggregation.
Int. J. Cancer
,
91
,
167
–172.

Tsuchiyama
,
Y.
, Wada, J., Zhang, H., Morita, Y., Hiragushi, K., Hida, K., Shikata, K., Yamamura, M., Kanwar, Y.S., and Makino, H. (
2000
) Efficacy of galectins in the amelioration of nephrotoxic serum nephritis in Wistar Kyoto rats.
Kidney Int
.,
58
,
1941
–1952.

Tsutsumi
,
T.
, Suzuki, T., Moriya, K., Shintani, Y., Fujie, H., Miyoshi, H., Matsuura, Y., Koike, K., and Miyamura, T. (
2003
) Hepatitis C virus core protein activates ERK and p38 MAPK in cooperation with ethanol in transgenic mice.
Hepatology
,
38
,
820
–828.

van den Brule
,
F.A.
, Buicu, C., Berchuck, A., Bast, R.C., Deprez, M., Liu, F.T., Cooper, D.N., Pieters, C., Sobel, M.E., and Castronovo, V. (
1996
) Expression of the, 67-kD laminin receptor, galectin-1, and galectin-3 in advanced human uterine adenocarcinoma.
Hum. Pathol
.,
27
,
1185
–1191.

van den Brule
,
F.
, Califice, S., Garnier, F., Fernandez, P.L., Berchuck, A., and Castronovo, V. (
2003
) Galectin-1 accumulation in the ovary carcinoma peritumoral stroma is induced by ovary carcinoma cells and affects both cancer cell proliferation and adhesion to laminin-1 and fibronectin.
Lab. Invest
.,
83
,
377
–386.

van den Brule
,
F.A.
, Fernandez, P.L., Buicu, C., Liu, F.T., Jackers, P., Lambotte, R., and Castronovo, V. (
1997
) Differential expression of galectin-1 and galectin-3 during first trimester human embryogenesis.
Dev. Dyn
.,
209
,
399
–405.

van den Brule
,
F.A.
, Waltregny, D., and Castronovo, V. (
2001
) Increased expression of galectin-1 in carcinoma-associated stroma predicts poor outcome in prostate carcinoma patients.
J. Pathol
.,
193
,
80
–87.

van der Leij
,
J.
, van den Berg, A., Blokzijl, T., Harms, G., van Goor, H., Zwiers, P., van Weeghel, R., Poppema, S., and Visser, L. (
2004
) Dimeric galectin-1 induces IL-10 production in T-lymphocytes: an important tool in the regulation of the immune response.
J. Pathol
.,
204
,
511
–518.

Vas
,
V.
, Fajka-Boja, R., Ion, G., Dudics, V., Monostori, E., and Uher, F. (
2005
) Biphasic effect of recombinant galectin-1 on the growth and death of early hematopoietic cells.
Stem Cells
,
23
,
279
–287.

Vicovac
,
L.
, Jankovic, M., and Cuperlovic, M. (
1998
) Galectin-1 and -3 in cells of the first trimester placental bed.
Hum. Reprod
.,
13
,
730
–735.

von Wolff
,
M.
, Wang, X., Gabius, H.J., and Strowitzki, T. (
2005
) Galectin fingerprinting in human endometrium and decidua during the menstrual cycle and in early gestation.
Mol. Hum. Reprod
.,
11
,
189
–194.

Vyakarnam
,
A.
, Dagher, S.F., Wang, J.L., and Patterson, R.J. (
1997
) Evidence for a role for galectin-1 in pre-mRNA splicing.
Mol. Cell. Biol
.,
17
,
4730
–4737.

Wada
,
M.
, Ono, S., Kadoya, T., Kawanami, T., Kurita, K., and Kato, T. (
2003
) Decreased galectin-1 immunoreactivity of the skin in amyotrophic lateral sclerosis.
J. Neurol. Sci
.,
208
,
67
–70.

Walzel
,
H.
, Blach, M., Hirabayashi, J., Kasai, K.I., and Brock, J. (
2000
) Involvement of CD2 and CD3 in galectin-1 induced signaling in human Jurkat T-cells.
Glycobiology
,
10
,
131
–140.

Walzel
,
H.
, Brock, J., Pohland, R., Vanselow, J., Tomek, W., Schneider, F., and Tiemann, U. (
2004
) Effects of galectin-1 on regulation of progesterone production in granulosa cells from pig ovaries in vitro.
Glycobiology
,
14
,
871
–881.

Walzel
,
H.
, Schulz, U., Neels, P., and Brock, J. (
1999
) Galectin-1, a natural ligand for the receptor-type protein tyrosine phosphatase CD45.
Immunol. Lett
.,
67
,
193
–202.

Wang
,
L.
, Friess, H., Zhu, Z., Frigeri, L., Zimmermann, A., Korc, M., Berberat, P.O., and Buchler, M.W. (
2000
) Galectin-1 and galectin-3 in chronic pancreatitis.
Lab. Invest
.,
80
,
1233
–1241.

Wang
,
P.
, Mariman, E., Keijer, J., Bouwman, F., Noben, J.P., Robben, J., and Renes, J. (
2004
) Profiling of the secreted proteins during 3T3-L1 adipocyte differentiation leads to the identification of novel adipokines.
Cell. Mol. Life Sci
.
61
,
2405
–2417.

Wasano
,
K.
and Hirakawa, Y. (
1997
) Recombinant galectin-1 recognizes mucin and epithelial cell surface glycocalyces of gastrointestinal tract.
J. Histochem. Cytochem
.,
45
,
275
–283.

Wells
,
V.
, Davies, D., and Mallucci, L. (
1999
) Cell cycle arrest and induction of apoptosis by beta galactoside binding protein (beta GBP) in human mammary cancer cells. A potential new approach to cancer control.
Eur. J. Cancer
,
35
,
978
–983.

Wollina
,
U.
, Schreiber, G., Gornig, M., Feldrappe, S., Burchert, M., and Gabius, H.J. (
1999
) Sertoli cell expression of galectin-1 and -3 and accessible binding sites in normal human testis and Sertoli cell only syndrome.
Histol. Histopathol
.,
14
,
779
–784.

Woynarowska
,
B.
, Skrincosky, D.M., Haag, A., Sharma, M., Matta, K., and Bernacki, R.J. (
1994
) Inhibition of lectin-mediated ovarian tumor cell adhesion by sugar analogs.
J. Biol. Chem
.,
269
,
22797
–22803.

Xu
,
Z.
and Weiss, A. (
2002
) Negative regulation of CD45 by differential homodimerization of the alternatively spliced isoforms.
Nat. Immunol
.,
3
,
764
–771.

Yamaoka
,
K.
, Mishima, K., Nagashima, Y., Asai, A., Sanai, Y., and Kirino, T. (
2000
) Expression of galectin-1 mRNA correlates with the malignant potential of human gliomas and expression of antisense galectin-1 inhibits the growth of 9 glioma cells.
J. Neurosci. Res
.,
59
,
722
–730.

Zou
,
W.
(
2005
) Immunosuppressive networks in the tumour environment and their therapeutic relevance.
Nat. Rev. Cancer
,
5
,
263
–274.

Zuniga
,
E.
, Gruppi, A., Hirabayashi, J., Kasai, K.I., and Rabinovich, G.A. (
2001
) Regulated expression and effect of galectin-1 on Trypanosoma cruzi-infected macrophages: modulation of microbicidal activity and survival.
Infect. Immun
.,
69
,
6804
–6812.

Author notes

2Laboratory of Toxicology, Institute of Pharmacy, Free University of Brussels (ULB), Brussels; 3XPeDoc sprl, rue Halvaux 37, 7090 Ronquières; and 4Department of Neurosurgery, Erasmus University Hospital, Free University of Brussels (ULB), Brussels, Belgium