Soluble vascular cell adhesion molecule-1 (sVCAM-1) is generated during inflammation and can alter lymphocyte functions. The authors report that the binding of sVCAM-1 to 4 integrin-bearing cells is a dynamically regulated, active cellular process. Binding of recombinant sVCAM-1 to 4 integrins on peripheral blood mononuclear cells was cell-type specific. Circulating CD16+ NK cells constitutively bound sVCAM-1 with high affinity, whereas a subpopulation of T-lymphocytes, primarily CD45RO+ (memory), bound sVCAM-1 only after phorbol ester stimulation. sVCAM-1 binding to homogenous stable cell lines was also cell-type specific, and required active cellular processes because it was blocked by the inhibition of ATP synthesis and by Fas-induced apoptosis. Indeed, the loss of high-affinity VCAM-1 binding was an early event in apoptosis. Furthermore, an H-Ras/Raf-initiated signaling pathway also suppressed sVCAM-1 binding to 4β1 integrins. Collectively, these results showed that the capacity of 4 integrins to bind VCAM-1 is actively regulated and that this regulation may control 4 integrin-dependent cellular functions.

The α4 integrins are adhesion receptors widely expressed on mononuclear leukocytes and eosinophils.1 A major ligand for α4 integrins is vascular cell adhesion molecule-1 (VCAM-1).2 VCAM-1 is a member of the immunoglobulin superfamily and is expressed on vascular endothelium at sites of inflammation and on selected other sites, such as thymic cortical epithelium and bone marrow stroma.2,3 Cells expressing α4 integrins can interact with VCAM-1 to mediate events such as the recruitment of leukocytes to sites of inflammation and the development and activation of lymphocytes.4,5 

In addition to its cell surface expression, VCAM-1 has been found in a soluble form in the culture supernatant of cytokine-treated endothelial cells, in human serum, and in the synovial fluid of patients with rheumatoid arthritis.6,7 The origin of this soluble VCAM-1 (sVCAM-1) is not entirely known, but release from the cell surface by proteolytic cleavage is possible.8 sVCAM-1 may act to suppress further leukocyte migration to tissues by removing the ligand for α4 integrins. The soluble form of VCAM-1 could also act as a competitive inhibitor of ligand binding. In addition, the interaction of the sVCAM-1 with α4 integrins could provide a signal to alter cell function. Indeed, in T cells, sVCAM-1 acts as a chemotractant,9 and sVCAM-1 binding inhibits T-cell activation.10 

Integrin function can be dynamically regulated without changes in integrin expression, a process termed “inside-out signaling.”11,12 Two general mechanisms account for such regulation. One involves a change in the affinity of ligand binding, operationally defined as “activation.”11Alternatively, affinity-independent mechanisms, such as changes in cell shape, spreading, and lateral mobility/clustering of integrins, also modulate cell adhesion.13-15 Many integrins, such as αIIbβ3, α5β1, and αMβ2, manifest active cellular regulation of the affinity of ligand binding.16,17 In sharp contrast, active cellular regulation of α4β1 affinity for soluble ligand is controversial, and a recent review emphasized that unlike other integrins, the affinity of α4β1 may not be physiologically regulated.13 It is clear that soluble ligand binding to α4β1 can be increased by exogenous agents such as Mn++ or activating antibodies.18 A report has demonstrated increased soluble fibronectin (FN) binding to α4β1 on T cells stimulated by L-selectin engagement, but its significance is unclear.19 Furthermore, several studies have implicated α4β1 activation indirectly with reported antibodies such as monoclonal antibody (mAb) 15/7, but whether these antibodies truly recognize high-affinity integrin conformations has been questioned.9,13,20,21 

In the current study, we investigated the regulation of sVCAM-1 binding to α4β1 integrin using a sVCAM-1-immunoglobulin fusion protein as ligand. We report that sVCAM-1 binding to α4 integrins was cell-type–specific and energy dependent and that it was disrupted during apoptosis. Specific intracellular signaling pathways also regulated sVCAM-1 binding to α4β1. Thus, the function of the α4β1 integrin can be physiologically regulated by changes in soluble ligand binding.

Antibodies and reagents

The antihuman β1 mAb, 8A2, was the generous gift of Drs N. Kovach and J. Harlan (University of Washington, Seattle, WA). Antihuman α4 (9F10), antihuman β1 (9EG7), and fluorescein isothiocyanate (FITC)-conjugated antihuman CD3 (HIT3a) mAbs were purchased from PharMingen (San Diego, CA). Function-blocking antihuman α4 (HP2/1) and the anti-Fas (CH11) IgM were purchased from Immunotech (Westbrook, ME). The antihuman CD3 (OKT3) hybridoma was obtained from American Type Culture Collection (ATCC; Rockville, MD) and IgG purified as previously described.17 Phorbol 12-myristate 13-acetate (PMA) and sodium azide were purchased from Sigma Chemical (St. Louis, MO), and 2-deoxyglucose was purchased from Calbiochem (La Jolla, CA). Anti-Tac antibody, 7G7B6, was obtained from ATCC, and biotinylated with biotin-N-hydroxy-succinimide was obtained from Sigma Chemical. Phycoerythrin (PE)-conjugated annexin V was purchased from R&D Systems (Minneapolis, MN). PE-conjugated streptavidin was purchased from Molecular Probes (Eugene, OR). A VCAM–Ig fusion protein containing the 2 N-terminal domains of human VCAM-1 fused to the human IgG1 heavy chain was a generous gift of Dr Roy Lobb (Biogen, Cambridge, MA) and was prepared as previously described.18 A VCAM-Cκ fusion protein containing the 7 Ig domains of human VCAM-1 fused to the murine Cκ segment was a gift from Dr. Richard M. Wright (Novartis Pharmaceuticals, Summit, NJ). Purified human plasma FN was purchased from Life Technologies (Grand Island, NY) and fluoresced as follows: FN was resuspended in phosphate-buffered saline at 5 mg/mL and was added to a reaction buffer containing 0.1 mol/L NaHCO3, pH 9, and 2 mg/mL FITC-celite (Sigma Chemical). The solution was incubated for 60 minutes at room temperature in the dark. The FITC-FN was purified on a PD-10 Sephadex column from AmershamPharmacia Biotech (Wikstroms, Sweden) and analyzed by spectrophotometry.

Cells

The T-leukemic cell lines Jurkat and HUT-78, the monocytic cell lines THP-1 and U937, and the rat basophilic leukemia (RBL) cell line were purchased from ATCC and cultured in RPMI-1640 medium (Biowhittaker, Walkersville, MD) supplemented with 10% fetal calf serum (Biowhittaker), 1% glutamine, 1% penicillin, and 1% streptomycin (Sigma Chemical). Human lymphocytes were purified from peripheral blood from normal donors by centrifugation through a Ficoll-Paque gradient (Pharmacia Biotech, Uppsala, Sweden), panning for monocytes on tissue culture plastic, and passaging over a nylon wool column.17 Chinese hamster ovary (CHO) cells expressing both human α4 and β1 integrin subunits were a generous gift from Dr Yoshikazu Takada (The Scripps Research Institute, La Jolla, CA) and have been described elsewhere.22 

cDNA constructs and transfections

The chimeric Tac-α5 and Tac-β1 constructs in CDM8 vector have been previously described.23 The pDCR-H-Ras (G12V) was a gift from Dr M. H. Wigler (Cold Spring Harbor Laboratory, Cold Spring Harbor, NY). The vector encoding RafBxB has been previously described.24 CHO-α4β1 cells were transiently transfected using lipofectamine (Gibco BRL, Grand Island, NY) according to the manufacturer's instructions, as previously described.25 

Construction and expression of VCAM-1-immunoglobulin fusion protein

Total RNA was isolated from HUVECs stimulated with IL-1β using Ultraspec (BioTecX, Houston, TX) following the instructions of the manufacturer. First-strand cDNA was generated using the GeneAmp kit of Perkin/Elmer (Norwalk, CT) and oligo-dT as a primer following the manufacturer's instructions. The entire coding region of the VCAM-1 cDNA was cloned to generate pRmHaVCAM. The entire coding sequence of the extracellular domain (7 Ig-like domains) of VCAM-1 was polymerase chain reaction-amplified from pRmHaVCAM using the following primers: 5′-GCTAGCGTCGACATGCCTGGGAAGATGGTC-3′; 5′-AGCGTCAGCCTCAGGAGAAAAATAGTC-3′. The resultant fragment was then subcloned into the NheI site of pBB4Ig, which contains the human Fc coding sequences at the carboxy terminus of the recombinant protein. Recombinant protein was expressed in Sf9 cells grown in SF900 II SFM (Life Technologies, Gaithersburg, MD). Recombinant protein was purified using Protein A filters (Nygene, Goldens Bridge, NY). Analysis of recombinant protein by sodium dodecyl sulphate–polyacrylamide gel electrophoresis revealed a major molecular species of ∼240 kd under nonreducing conditions and ∼115 kd under reducing conditions. These are the predicted weights of the dimer and the monomer, respectively.

Soluble ligand-binding assay

Cells (5 × 105) were resuspended in Dulbecco's minimum essential medium (DMEM; Biowhittaker) containing 0.1% bovine serum albumin (BSA; Sigma Chemical) or in a modified Tyrode's buffer (150 mmol/L NaCl, 2.5 mmol/L KCl, 12 mmol/L NaHCO3, 1 mg/mL glucose, and 1 mg/mL BSA) with 1 mmol/L MnCl2. A modified Tyrode's buffer was used with MnCl2 because this divalent cation precipitates in phosphate-containing solutions. The VCAM-Ig fusion protein, or FITC-FN, was added to the mixture and incubated for 30 minutes at room temperature (preliminary studies indicated no significant differences in ligand binding at room temperature and 37°C). Afterward, cells were washed twice with 0.5 mL of their respective buffers. If cells were also being analyzed for surface antigen expression, they were incubated with primary antibody for 30 minutes on ice and again washed twice. Cells were then resuspended in the same buffer containing either PE- or FITC-conjugated donkey antihuman IgG (Jackson Immunoresearch, West Grove, PA) at a 1:100 dilution (if they were dual labeled, FITC-conjugated antimouse IgG was also added) and incubated for 30 minutes. Cells were washed twice and fluorescence analyzed on a FACSCalibur flow cytometer (Becton Dickinson, Mountain View, CA) using CellQuest software.

Analysis of apoptosis

The assessment of apoptosis was measured by the binding of PE-conjugated annexin V to externalized phosphatidylserine by flow cytometry according to the manufacturer's instructions (R&D Systems).

sVCAM-1 binding to isolated human peripheral blood cells

Because the interaction of α4 integrins with VCAM-1 modulates the functions of mononuclear leukocytes, we examined the regulation of sVCAM-1 binding to human peripheral blood lymphoid cells. These cells were analyzed for both CD3 expression and sVCAM-1 (7 Ig-domain form) binding by 2-color flow cytometry. A population of CD3-negative cells bound sVCAM-1, whereas few (typically less than 2%) of the CD3-positive T-lymphocytes manifested such constitutive binding (Figure1) The α4 integrins were functional on these CD3+ T-lymphocytes because 60% of these cells bound sVCAM-1 in the presence of an exogenous integrin activator, MnCl2 (data not shown). Stimulation with PMA (50 ng/mL) for 30 minutes resulted in increased sVCAM-1 binding in a subgroup representing ∼15% of the CD3-positive T-lymphocytes (Figure 1). Binding in each case was α4 integrin specific because it was completely blocked by an anti-α4 antibody, HP2/1. Furthermore, binding was not simply ascribable to the dimeric nature of the VCAM-Ig construct. A monomeric VCAM-1 construct containing the 7 extracellular Ig domains fused to murine κ light chain showed similar binding characteristics (data not shown). These results indicated that a subgroup of peripheral T-lymphocytes responds to PMA stimulation with increased sVCAM-1 binding to α4 integrins.

Fig. 1.

sVCAM-1 binding to isolated peripheral blood lymphocytes.

Human peripheral blood lymphoid cells were isolated as described in “Materials and Methods.” Cells were resuspended in Tyrode's buffer containing 1 mmol/L CaCl2, 1 mmol/L MgCl2, and 0.1% BSA and incubated for 30 minutes at room temperature in the absence of stimulation (upper panels) or in the presence of PMA (50 ng/mL) (lower panels). In the presence or absence of the function blocking anti-α4 mAb HP2/1, 7 Ig-domain VCAM-Ig (20 μg/mL) was added to the cells, and the mixture was incubated for 30 minutes at room temperature. After washing, the cells were stained with FITC-anti-CD3 mAb (HIT3a). At the same time, they were stained with PE-conjugated donkey antihuman IgG to detect the VCAM-Ig fusion protein as described in “Materials and Methods.” The percentage of cells in quadrants is indicated. Results are representative of 3 separate experiments.

Fig. 1.

sVCAM-1 binding to isolated peripheral blood lymphocytes.

Human peripheral blood lymphoid cells were isolated as described in “Materials and Methods.” Cells were resuspended in Tyrode's buffer containing 1 mmol/L CaCl2, 1 mmol/L MgCl2, and 0.1% BSA and incubated for 30 minutes at room temperature in the absence of stimulation (upper panels) or in the presence of PMA (50 ng/mL) (lower panels). In the presence or absence of the function blocking anti-α4 mAb HP2/1, 7 Ig-domain VCAM-Ig (20 μg/mL) was added to the cells, and the mixture was incubated for 30 minutes at room temperature. After washing, the cells were stained with FITC-anti-CD3 mAb (HIT3a). At the same time, they were stained with PE-conjugated donkey antihuman IgG to detect the VCAM-Ig fusion protein as described in “Materials and Methods.” The percentage of cells in quadrants is indicated. Results are representative of 3 separate experiments.

Close modal

We next used surface markers to characterize the subpopulation of CD3-positive T-lymphocytes that bound sVCAM-1 in response to PMA stimulation. Resting CD45RO+ (memory) and CD45RA+ (naive) T-lymphocytes showed little difference in sVCAM-1 binding (Figure2). However, after PMA stimulation, CD45RO+ cells bound much more sVCAM-1 than CD45RA+ cells (Figure 2). This CD45RO+ cell response was not simply a response to the differences in α4-integrin expression because sVCAM-1 binding could be induced on both CD45RO+ and CD45RA+ cells after exogenous activation with MnCl2 (data not shown). These results showed that CD45RO+ memory T-lymphocytes responded to phorbol ester stimulation by increasing soluble VCAM-1 binding.

Fig. 2.

PMA stimulates sVCAM-1 binding to CD45RO-positive peripheral T-lymphocytes.

Human peripheral blood T-lymphocytes were isolated as described in “Materials and Methods”. Cells were resuspended in Tyrode's buffer containing 1 mmol/L CaCl2, 1 mmol/L MgCl2, and 0.1% BSA (solid line) or 5 mmol/L EDTA (shaded line). They were incubated for 30 minutes at room temperature in the absence of stimulation (upper panels) or in the presence of PMA (50 ng/ml) (lower panels). 7 Ig-domain VCAM-Ig (20 μg/mL) was added, and binding was allowed to proceed for 30 minutes at room temperature. After they were washed, cells were incubated with anti-CD45RA mAb (HI100) or anti-CD45RO mAb (UCHL1). Binding of VCAM-Ig and anti-CD45 mAbs was detected with PE-conjugated donkey antihuman IgG and FITC-conjugated donkey antimouse IgG, respectively, as described in “Materials and Methods.” Histograms depict sVCAM-1 binding to cells gated for CD45 expression. To obtain a quantitative estimate of sVCAM-1 binding, we calculated a numeric activation index (AI) defined as 100 × (Fo−Fr)/Fmax−Fr), where Fo is the mean fluorescence intensity of sVCAM-1 binding, Fr is background fluorescence in the presence of 5 mmol/L EDTA, and Fmax is the fluorescence intensity in the presence of 1 mmol/L MnCl2 (not shown). Results are representative of 3 separate experiments.

Fig. 2.

PMA stimulates sVCAM-1 binding to CD45RO-positive peripheral T-lymphocytes.

Human peripheral blood T-lymphocytes were isolated as described in “Materials and Methods”. Cells were resuspended in Tyrode's buffer containing 1 mmol/L CaCl2, 1 mmol/L MgCl2, and 0.1% BSA (solid line) or 5 mmol/L EDTA (shaded line). They were incubated for 30 minutes at room temperature in the absence of stimulation (upper panels) or in the presence of PMA (50 ng/ml) (lower panels). 7 Ig-domain VCAM-Ig (20 μg/mL) was added, and binding was allowed to proceed for 30 minutes at room temperature. After they were washed, cells were incubated with anti-CD45RA mAb (HI100) or anti-CD45RO mAb (UCHL1). Binding of VCAM-Ig and anti-CD45 mAbs was detected with PE-conjugated donkey antihuman IgG and FITC-conjugated donkey antimouse IgG, respectively, as described in “Materials and Methods.” Histograms depict sVCAM-1 binding to cells gated for CD45 expression. To obtain a quantitative estimate of sVCAM-1 binding, we calculated a numeric activation index (AI) defined as 100 × (Fo−Fr)/Fmax−Fr), where Fo is the mean fluorescence intensity of sVCAM-1 binding, Fr is background fluorescence in the presence of 5 mmol/L EDTA, and Fmax is the fluorescence intensity in the presence of 1 mmol/L MnCl2 (not shown). Results are representative of 3 separate experiments.

Close modal

We also characterized the CD3-negative cells that manifested constitutive VCAM-1 binding. This population of cells had the size and granularity of lymphoid cells in flow cytometry. They were not, however, depleted by passage through a nylon wool column, suggesting that they were not B-lymphocytes. Using CD16 as a surface marker of natural killer (NK) cells, we found that more than 50% of circulating CD16+ NK cells constitutively bound sVCAM-1 and that this binding was inhibited by the function-blocking anti-α4 antibody, HP2/1 (Figure 3). Thus, most circulating CD16+ NK cells express α4 integrins that constitutively bind soluble VCAM-1. Analysis of sVCAM-1 binding to circulating B cells was not possible because of the interaction of our secondary detection antibody with surface-bound IgG or Fc-receptors (data not shown).

Fig. 3.

CD16-positive human peripheral blood natural killer cells constitutively bind sVCAM-1.

Human peripheral blood lymphoid cells were isolated as described in “Materials and Methods”. Cells were resuspended in Tyrode's buffer containing 1 mmol/L, CaCl2, 1 mmol/L MgCl2, and 0.1% BSA. In the presence or absence of the function blocking anti-α4 mAb HP2/1 (10 μg/mL), 7 Ig-domainVCAM-Ig (20 μg/mL) was added, and cells were incubated for 30 minutes at room temperature. After they were washed, cells were incubated with FITC-anti-CD16 mAb.3G8 Binding of VCAM-Ig was detected with PE-conjugated donkey antihuman IgG as described in “Materials and Methods.” The percentage of cells in quadrants is indicated. Results are representative of 3 separate experiments.

Fig. 3.

CD16-positive human peripheral blood natural killer cells constitutively bind sVCAM-1.

Human peripheral blood lymphoid cells were isolated as described in “Materials and Methods”. Cells were resuspended in Tyrode's buffer containing 1 mmol/L, CaCl2, 1 mmol/L MgCl2, and 0.1% BSA. In the presence or absence of the function blocking anti-α4 mAb HP2/1 (10 μg/mL), 7 Ig-domainVCAM-Ig (20 μg/mL) was added, and cells were incubated for 30 minutes at room temperature. After they were washed, cells were incubated with FITC-anti-CD16 mAb.3G8 Binding of VCAM-Ig was detected with PE-conjugated donkey antihuman IgG as described in “Materials and Methods.” The percentage of cells in quadrants is indicated. Results are representative of 3 separate experiments.

Close modal

Cell-type–specific binding of sVCAM-1 to 4β1 integrin

To assess sVCAM-1 binding to homogenous cell populations, we examined binding to stable cell lines. In the Jurkat T-leukemia cell line, sVCAM-1 (7 Ig domain) bound tightly (EC50 ∼50 nmol/L), and binding was completely blocked by the anti-α4 antibody HP2/1 (Figure 4a) and by the anti-VCAM-1 antibody P8B1(data not shown). The addition of 1 mmol/L MnCl2, to activate α4β1 integrins exogenously resulted in ∼2-fold greater maximal sVCAM-1 binding at saturation. Thus, ∼50% of the functionally expressed α4β1 integrins constitutively bound sVCAM-1. Similar results were obtained with α4β1-transfected CHO cells (CHO-α4β1) (Figure 5). In contrast to sVCAM-1 binding, soluble FN showed no detectable binding to Jurkat cells unless integrins were exogenously activated with MnCl2 (Figure 4B). More than 80% of this binding was inhibited by HP2/1, reflecting the 5-fold greater abundance of α4β1 than other FN binding integrins on these cells (data not shown). Furthermore, the Mn++-induced soluble FN binding was inhibited by antibodies hfn7.1 and FN1-8 (data not shown), which recognize the ninth and tenth type 2I homology segments of FN.26,27 These results suggest that this binding occurs through the central cell-binding region of FN, as has been reported.28 In addition, minimal sVCAM-1 binding to α4β1 was seen with a VCAM-Ig construct containing only the 2 N-terminal Ig-like domains unless cells were exogenously activated with 1 mmol/L MnCl2 (Figure 4c).

Fig. 4.

High affinity binding of a sVCAM-Ig fusion protein and soluble FN to 4β1 integrins on Jurkat cells.

Cells were resuspended in DMEM containing 0.1% BSA or in a Tyrode's buffer containing 1 mmol/L MnCl2 and 0.1% BSA (Mn++). The indicated concentrations of VCAM-Ig (A) or FITC-FN (B) were added, and the mixtures were incubated for 30 minutes at room temperature. ○, Mn++ added; □, no addition. Where indicated by • and ▪, function-blocking anti-α4 mAb HP2/1 was added before the soluble ligand. VCAM-Ig binding was evaluated with FITC-conjugated donkey antihuman IgG using flow cytometry as described in “Materials and Methods” and expressed as geometric mean fluorescence intensity (MFI). (C) Jurkat cells were resuspended in either DMEM containing 0.1% BSA (solid line), DMEM with the addition of 5 mmol/L EDTA (shaded line), or Tyrode's buffer containing 1 mmol/L MnCl2(dotted line). VCAM-Ig containing only the first 2 N-terminal or all 7 Ig domains (20 μg/mL) was added to the cells, and the mixtures were incubated at room temperature for 30 minutes. Results are representative of 3 separate experiments.

Fig. 4.

High affinity binding of a sVCAM-Ig fusion protein and soluble FN to 4β1 integrins on Jurkat cells.

Cells were resuspended in DMEM containing 0.1% BSA or in a Tyrode's buffer containing 1 mmol/L MnCl2 and 0.1% BSA (Mn++). The indicated concentrations of VCAM-Ig (A) or FITC-FN (B) were added, and the mixtures were incubated for 30 minutes at room temperature. ○, Mn++ added; □, no addition. Where indicated by • and ▪, function-blocking anti-α4 mAb HP2/1 was added before the soluble ligand. VCAM-Ig binding was evaluated with FITC-conjugated donkey antihuman IgG using flow cytometry as described in “Materials and Methods” and expressed as geometric mean fluorescence intensity (MFI). (C) Jurkat cells were resuspended in either DMEM containing 0.1% BSA (solid line), DMEM with the addition of 5 mmol/L EDTA (shaded line), or Tyrode's buffer containing 1 mmol/L MnCl2(dotted line). VCAM-Ig containing only the first 2 N-terminal or all 7 Ig domains (20 μg/mL) was added to the cells, and the mixtures were incubated at room temperature for 30 minutes. Results are representative of 3 separate experiments.

Close modal
Fig. 5.

Cell-type–specific binding of sVCAM-1 to 4β1 integrin.

Jurkat, CHO-α4β1, RBL, THP-1, U937, or HUT-78 was resuspended in either DMEM containing 0.1% BSA (solid line), DMEM with the addition of 5 mmol/L EDTA (shaded line), or Tyrode's buffer containing 1 mmol/L MnCl2 (dotted line). Seven Ig-domain VCAM-Ig (20 μg/mL) was added, and the mixture was incubated at room temperature for 30 minutes. VCAM-Ig binding was detected with FITC-conjugated donkey antihuman IgG as described in “Materials and Methods.” Results are representative of 3 separate experiments.

Fig. 5.

Cell-type–specific binding of sVCAM-1 to 4β1 integrin.

Jurkat, CHO-α4β1, RBL, THP-1, U937, or HUT-78 was resuspended in either DMEM containing 0.1% BSA (solid line), DMEM with the addition of 5 mmol/L EDTA (shaded line), or Tyrode's buffer containing 1 mmol/L MnCl2 (dotted line). Seven Ig-domain VCAM-Ig (20 μg/mL) was added, and the mixture was incubated at room temperature for 30 minutes. VCAM-Ig binding was detected with FITC-conjugated donkey antihuman IgG as described in “Materials and Methods.” Results are representative of 3 separate experiments.

Close modal

To assess further the cell-type specificity of sVCAM-1 binding (the 7 Ig domain form), we assayed other α4 expressing cell lines. In contrast to Jurkat cells, the human T-leukemic line HUT-78 bound sVCAM-1 minimally (Figure 5). However, the α4β1 on these cells was functional because soluble VCAM-Ig binding could be induced either by treatment with MnCl2 or by the β1 integrin activating antibodies 8A2 and 9EG7 (data not shown). Differences in sVCAM-1 binding were not attributable to differences in α4β1 expression because they express similar integrin levels, reflected by the similar sVCAM-1 binding in the presence of MnCl2. The human monocytic lines THP-1 and U937 also bound little sVCAM-1 without exogenous activation, whereas RBL bound sVCAM-1 in a manner similar to Jurkat and CHO-α4β1 cells (Figure 5). These results indicated that the capacity of α4β1 integrin to bind sVCAM-1 is dependent on the cellular context.

sVCAM-1 binding to 4β1 integrin is dependent on active cellular processes

We next questioned whether binding of sVCAM-1 to α4β1 is a passive or an active cellular event. Treating Jurkat and CHO-α4β1 cells with the metabolic inhibitors sodium azide and 2-deoxyglucose resulted in a time-dependent suppression of sVCAM-1 binding (Figure6). This decrease in sVCAM-1 binding was not caused by a loss of α4β1 integrin expression because the binding could be recovered by exogenous activation with Mn++. Thus, metabolic energy is required to maintain sVCAM-Ig binding to α4β1 integrin.

Fig. 6.

Cellular-energy depletion suppresses sVCAM-Ig binding.

Jurkat (A) or CHO-α4β1 (B) cells were placed in glucose-free Tyrode's buffer containing 2-deoxyglucose (2 mg/mL) and sodium azide (0.1% wt/vol) and were incubated at 37°C for 0, 30, 60, or 120 minutes. Subsequently, cells were resuspended in DMEM or Tyrode's buffer containing 1 mmol/L MnCl2 (Mn++), and 7 Ig-domain VCAM-Ig binding was assessed as described in “Materials and Methods”. VCAM-Ig binding was expressed as a percentage of initial binding in cells incubated in the absence of 2-deoxyglucose and sodium azide. Results are the mean ± SEM of 3 independent experiments. ◊, Mn++ added; □, no addition.

Fig. 6.

Cellular-energy depletion suppresses sVCAM-Ig binding.

Jurkat (A) or CHO-α4β1 (B) cells were placed in glucose-free Tyrode's buffer containing 2-deoxyglucose (2 mg/mL) and sodium azide (0.1% wt/vol) and were incubated at 37°C for 0, 30, 60, or 120 minutes. Subsequently, cells were resuspended in DMEM or Tyrode's buffer containing 1 mmol/L MnCl2 (Mn++), and 7 Ig-domain VCAM-Ig binding was assessed as described in “Materials and Methods”. VCAM-Ig binding was expressed as a percentage of initial binding in cells incubated in the absence of 2-deoxyglucose and sodium azide. Results are the mean ± SEM of 3 independent experiments. ◊, Mn++ added; □, no addition.

Close modal

sVCAM-1 binding is disrupted during apoptosis

We next determined whether a physiologic relevant stimulus could modulate the activation of α4β1. An early feature of apoptotic cell death is loss of cell adhesion. Thus, we examined the effect of Fas-induced apoptotic cell death on sVCAM-1 binding to Jurkat cells. Treatment of Jurkat cells with anti-Fas antibody resulted in apoptosis, as measured by annexin-V binding to surface-expressed phosphatidylserine (Figure 7). The annexin-V positive cells (apoptotic) manifested markedly reduced sVCAM-1 binding compared with annexin-V–negative cells (Figure 7B). However, the surface expression of α4 on apoptotic cells was not markedly different from that on nonapoptotic cells (Figure 7D). Furthermore, binding to apoptotic cells could be restored by adding the activating antibody 8A2 (Figure 7C). In addition, 8A2-induced sVCAM-1 binding increased annexin-V binding to the previously negative cells (Figure 7C). This suggests that sVCAM-1 binding promotes Fas-induced apoptosis in these cells. Through kinetic analysis, we found that the reduction in sVCAM-1 binding occurred rapidly after Fas cross-linking. A large proportion of cells lost sVCAM-1 binding while they remained annexin-V negative during the first 30 minutes after Fas ligation (Figure 7E). These results indicated that the disruption of sVCAM-1 binding is an early event in apoptosis.

Fig. 7.

Fas-induced apoptosis suppresses sVCAM-Ig binding to Jurkat cells.

Jurkat cells were untreated (A) or were treated with anti-Fas antibody CH11 (300 ng/mL) for 4 hours (B-D). 7 Ig-domain VCAM-Ig binding was assayed as described in “Materials and Methods” (A-C). (D) Cells were stained with the anti-α4-integrin mAb 9F10. Outer membrane phosphatidylserine exposure was assayed by PE-conjugated annexin-V binding, as described in “Materials and Methods”. The percentage of cells in each quadrant is indicated. Note the marked reduction in VCAM-Ig binding to annexin-V(+) cells (B) without a major reduction in α4 expression (D). VCAM-Ig binding to annexin-V(+) cells is restored by the activating antibody 8A2 (C). 8A2 also markedly increased annexin-V binding to previously negative cells (C). (E) Kinetics of sVCAM-1 and annexin-V binding to Jurkat cells after anti-Fas antibody treatment was analyzed. Cells that bound VCAM-Ig (V+) or failed to bind VCAM-Ig (V−) are indicated. Similarly, annexin-V binding (A+) and nonbinding (A−) cells are also indicated. Note that the A(−)/V(−) cells initially increased coordinately with the decrease in A(−)/V(+) cells, before any increase in A(+) cells, and note the nearly complete absence of V(+)/A(+) cells, indicating that in apoptotic cells α4-integrin activation is blocked. Results are representative of 3 separate experiments.

Fig. 7.

Fas-induced apoptosis suppresses sVCAM-Ig binding to Jurkat cells.

Jurkat cells were untreated (A) or were treated with anti-Fas antibody CH11 (300 ng/mL) for 4 hours (B-D). 7 Ig-domain VCAM-Ig binding was assayed as described in “Materials and Methods” (A-C). (D) Cells were stained with the anti-α4-integrin mAb 9F10. Outer membrane phosphatidylserine exposure was assayed by PE-conjugated annexin-V binding, as described in “Materials and Methods”. The percentage of cells in each quadrant is indicated. Note the marked reduction in VCAM-Ig binding to annexin-V(+) cells (B) without a major reduction in α4 expression (D). VCAM-Ig binding to annexin-V(+) cells is restored by the activating antibody 8A2 (C). 8A2 also markedly increased annexin-V binding to previously negative cells (C). (E) Kinetics of sVCAM-1 and annexin-V binding to Jurkat cells after anti-Fas antibody treatment was analyzed. Cells that bound VCAM-Ig (V+) or failed to bind VCAM-Ig (V−) are indicated. Similarly, annexin-V binding (A+) and nonbinding (A−) cells are also indicated. Note that the A(−)/V(−) cells initially increased coordinately with the decrease in A(−)/V(+) cells, before any increase in A(+) cells, and note the nearly complete absence of V(+)/A(+) cells, indicating that in apoptotic cells α4-integrin activation is blocked. Results are representative of 3 separate experiments.

Close modal

Suppression of 4β1-mediated sVCAM-1 binding by a Ras/Raf-initiated signaling pathway

A Ras/Raf-initiated signaling pathway suppressed the activation of chimeric integrins.25 Thus, we questioned whether such a signaling pathway could regulate α4β1-mediated soluble VCAM-1 binding. CHO-α4β1 cells were transiently cotransfected with cDNA encoding constitutively active H-Ras (H-RasG12V) or Raf-1 (RafBxB) and Tac-α5 (used as a surface marker of transfection). Forty-eight hours after transfection, cells were analyzed for Tac expression and VCAM-Ig binding by flow cytometry. Transfection of Tac-α5 alone, or in combination with vector control DNA, resulted in little change in sVCAM-1 binding (Figure8). Transfection with activated forms of H-Ras or Raf resulted in an approximately 50% reduction in sVCAM-1 binding (Figure 8), similar to the suppression seen by the over-expression of free β1 cytoplasmic domain (Tac-β1) (Figure 8), which has also been described for chimeric integrins.29This suppression was cell-autonomous; only transfected cells showed reduced sVCAM-1 binding. In addition, this reduced binding was not caused by a loss of α4β1 integrins given that activating the integrins with mAb 9EG7 (data not shown) could restore the binding. sVCAM-1 binding was also suppressed by the overexpression of wild-type H-Ras, but dominant negative H-Ras N17 and constitutively active CDC42 and Rac-1 had no effect on soluble ligand binding (data not shown). Thus, as with chimeric integrins, the binding of soluble ligands to α4β1 integrin is regulated by a Ras/Raf-mediated suppression pathway.

Fig. 8.

Activated H-Ras and Raf-1 suppress 4β1 integrin-mediated VCAM-1 ligand binding.

CHO-α4β1 cells were transiently transfected with cDNA encoding Tac-α5 or Tac-β1 alone or with a combination of Tac-α5 plus vector control DNA, H-Ras G12V, or RafBxB. After 48 hours, cells were analyzed for Tac expression and soluble VCAM-Ig binding by 2-color flow cytometry. 7 Ig-domain sVCAM-Ig binding was evaluated in low Tac-expressing cells, designated Tac(−) (□), and high Tac expressing cells, designated Tac(+) (▪). Percentage inhibition of VCAM-Ig binding was calculated relative to cells transfected with Tac-α5 cDNA alone. Results are the means ± SEM of 3 separate experiments.

Fig. 8.

Activated H-Ras and Raf-1 suppress 4β1 integrin-mediated VCAM-1 ligand binding.

CHO-α4β1 cells were transiently transfected with cDNA encoding Tac-α5 or Tac-β1 alone or with a combination of Tac-α5 plus vector control DNA, H-Ras G12V, or RafBxB. After 48 hours, cells were analyzed for Tac expression and soluble VCAM-Ig binding by 2-color flow cytometry. 7 Ig-domain sVCAM-Ig binding was evaluated in low Tac-expressing cells, designated Tac(−) (□), and high Tac expressing cells, designated Tac(+) (▪). Percentage inhibition of VCAM-Ig binding was calculated relative to cells transfected with Tac-α5 cDNA alone. Results are the means ± SEM of 3 separate experiments.

Close modal

Soluble VCAM-1 is generated at sites of inflammation and regulates lymphocyte function.9,10 We now report that active cellular processes regulate the binding of sVCAM-1 to the integrin α4β1. Binding of sVCAM-1 to α4β1 is cell-type specific in primary cells and in cell lines. In addition, binding is energy dependent and is disrupted during apoptosis. sVCAM-1 binding is also suppressed by constitutively active variants of Ras and Raf-1. CD45RO+ peripheral blood T cells are stimulated to bind sVCAM-1 by treatment with phorbol ester. Thus, the affinity of α4β1 for its ligand, VCAM-1, is regulated by active cellular events. Inside-out signaling through α4 integrins may regulate events such as leukocyte trafficking, migration, survival, and hematopoiesis.

Activation-dependent binding of soluble ligands to α4β1 integrin is differentially regulated. We found that a soluble recombinant fusion protein containing the complete extracellular 7 Ig domains of VCAM-1 bound spontaneously to α4β1 on Jurkat cells. In contrast, soluble FN failed to bind to these cells unless the integrins were exogenously activated, as has been described.19 Furthermore, a soluble VCAM-Ig protein consisting of the first 2 N-terminal Ig domains was a poor soluble ligand for α4β1 compared with the 7 Ig domain form. Ig domains 1 and 4 of VCAM-1 can mediate α4β1-dependent cell adhesion independently.30 Thus, it is unclear whether domain 4 is solely responsible for the active binding of sVCAM-1 or whether there is cooperation between domains 1 and 4. Mutational analysis of each of these domains is ongoing and will provide insight into this question. The binding observed in this study did not result simply from the dimeric nature of the fusion protein; a monomeric 7 domain VCAM-1 showed similar binding characteristics. However, cooperative interaction between domains 1 and 4 on a single VCAM-1 molecule with 1 or more α4β1 integrins could play a role in soluble ligand binding. Interestingly, a splice variant of VCAM-1 lacking domain 4 has been described.31 It will be of interest to determine whether differential expression of the 2 VCAM-1 variants and their interaction with α4β1 integrin may regulate the migration or development of α4 integrin-expressing cells.

The cellular context in which the α4β1 is expressed controls sVCAM-1 binding. Cell-type differences in adhesion to immobilized VCAM-1 correlated with sVCAM-1 binding. In general, cell types previously reported to adhere avidly to VCAM-1 (ie, Jurkat and CHO-α4β1) also showed the greatest sVCAM-1 binding. In contrast, cell types that adhere poorly to VCAM-1 (ie, THP-1 and HUT-78) bound less soluble ligand.20,32 In addition, the binding of sVCAM-1 to α4β1 integrin is an energy-dependent process because metabolic inhibitors suppressed soluble binding. Other investigators have failed to observe activation-dependent binding of a 7 Ig domain sVCAM-Ig fusion protein to α4 integrin-expressing cells.33,34 However, some of these studies were performed on a single cell type such as the erythroleukemic line K562,33 a cell line previously reported to express β1 integrins in a low-affinity state.35 Others failed to observe active sVCAM-1 binding to α4β1 at 4 °C.34Because we found that VCAM-1 binding is energy dependent, the process may have been blocked in experiments conducted at 4°C. Thus, sVCAM-1 binding is active and cell type-specific.

The differences in sVCAM-1 binding properties of α4 integrins on subgroups of blood mononuclear cells suggest important variations in α4 integrin-dependent functions. Many circulating CD16+ NK cells express α4 integrins in a high-affinity state based on constitutive sVCAM-1 binding. NK cells act as a first line of defense against many pathogens and accumulate rapidly in tissues early in the course of inflammation.36,37 The high-activation state of α4 integrins on NK cells may facilitate the attachment to endothelium early in inflammation, when VCAM-1 may be expressed at relatively low levels. Furthermore, high integrin affinity promotes cell migration at low-substrate densities.38 Consequently, NK cells may preferentially migrate at sites of low VCAM-1 expression. In contrast to NK cells, most circulating CD3-positive T-lymphocytes express α4 integrins that fail to bind sVCAM-1. Cellular activation may be required for the expression of high-affinity α4 integrins on these cells. PMA stimulation activated the α4 integrins, primarily on CD45RO+ memory T-lymphocytes rather than on CD45RA+ naı̈ve T cells. Future studies will determine whether more physiologically relevant stimuli, such as chemokines and T-cell-antigen-receptor engagement, will also activate α4 integrins on this T-cell subset. The activation of α4 integrins on these cells may facilitate their interaction with VCAM-1 and their recruitment to sites of inflammation. This is consistent with a paradigm of lymphocyte recirculation in which memory, but not naı̈ve, T-lymphocytes preferentially recirculate through peripheral blood vessels at inflammatory sites at which VCAM-1 is expressed.39,40 sVCAM-1 exhibits chemotactic activity for CD45RO+ memory T cells.9 This, too, may facilitate the influx of memory T cells to sites of inflammation. Interestingly, Giblin et al19 reported preferential activation of α4β1 and FN binding on naive T cells stimulated by L-selectins. Thus, the differential activation of α4β1 and ligand preference on naive and memory T cells may facilitate the recruitment of each of these T-cell subsets to selective sites.

A soluble form of VCAM-1 is present in the blood and synovial fluid of patients with inflammatory diseases at concentrations of 10 to 20 nmol/L,9 and higher levels could occur at localized regions of inflammation. The ED50 of ∼50 nmol/L for the binding of sVCAM-1 to Jurkat cells indicates that the pathologic levels of sVCAM-1 could occupy a significant fraction of activated α4β1. Perhaps sVCAM-1 acts as a competitive ligand inhibitor blocking adhesion to inflamed vascular endothelium. Furthermore, Kitani et al10 have shown that sVCAM-1 inhibits T-cell proliferation possibly because of the suppression of IL-2 production. Recent studies suggest that sVCAM-1 binding to T cells may induce apoptotic cell death (Ishii JK, et al, manuscript submitted for publication). Hence, though the expression of VCAM-1 on vascular endothelium is involved in the recruitment of α4-expressing cells to sites of inflammation, sVCAM-1 may dampen or downregulate the inflammatory response by altering leukocyte trafficking, activation, and survival.

Specific intracellular signaling pathways regulate the binding of sVCAM-1 to α 4 β 1 integrins. We observed that activated variants of H-Ras or Raf-1 inhibited α4β1 ligand binding. Thus, the activation state of α4β1, like chimeric αIIbβ3,25is subject to downregulation by an integrin-suppression pathway. Suppression of VCAM-1 binding to α4β1 could play a role in the release of mononuclear cells from sites such as bone marrow and thymus. For example, double-positive thymocytes express α4β1 in a constitutively active state that facilitates adhesion to VCAM-1 expressed in the cortex and at the corticomedullary junction of the thymus.3,41 A decrease in α4β1 affinity for VCAM-1 could be important in allowing the release of thymocytes from the cortex, facilitating their migration to the medulla. We are now in a position to study how extracellular stimuli, acting through Ras/Raf signaling, modulate α4β1 affinity.

sVCAM-1 binding to α4β1 on Jurkat cells was disrupted early in Fas-induced apoptosis. This disruption was not associated with a significant loss of α4-integrin expression, suggesting that the affinity of α4β1 for VCAM-1 was reduced. Cell detachment from the substratum and neighboring cells is a hallmark feature of apoptosis.42,43 How this occurs is unknown, but caspase-dependent proteolysis of proteins such as actin, fodrin, β-catenin, and FAK could play a role in this process.44-46 Our results suggest that decreased integrin affinity for ligands may play a role in the loss of cell adhesion associated with apoptosis. Apoptosis could suppress VCAM-1 binding by the induction of a suppressor-signaling pathway or by the proteolytic cleavage of integrin activators or of integrin cytoplasmic domains themselves.47,48 Our results now make it possible to analyze this newly observed early event in T-cell apoptosis.

Integrin-affinity modulation plays several roles in cellular processes, such as adhesion, migration, and extracellular matrix assembly. The prototypic integrin in which this form of regulation has been studied is platelet αIIbβ3. Using recombinant αIIbβ3, a number of novel integrin-regulatory molecules have been studied by genetic and biochemical approaches.25,29,49 However, there are cell-type and integrin-specific differences in integrin regulation. The use of soluble VCAM-1 offers a useful means to study affinity modulation of α4 integrins expressed in their natural cellular environment. Thus, genetic strategies, such as those developed for αIIbβ3, can now be used to dissect the regulation of α4β1 function in mononuclear cells. sVCAM-1 itself appears to be a ligand that regulates mononuclear cell function. Thus, the alteration of VCAM-1 binding through α4-integrin activation is likely to contribute to mononuclear cell-mediated inflammation and immune responses.

The authors thank Richard M. Wright (Novartis Pharmaceuticals, Summit, NJ) for the VCAM-Cκ fusion protein and Roy Lobb (Biogen, Cambridge, MA) for the 2-domain VCAM-Ig.

Supported by grants from the National Institutes of Health and by grant 3FB-0164 from the California Breast Cancer Research Program of the University of California.

Reprints:Mark H. Ginsberg, Department of Vascular Biology, The Scripps Research Institute, 10550 N Torrey Pines Rd, VB-2, La Jolla, CA 92037; e-mail: ginsberg@scripps.edu.

The publication costs of this article were defrayed in part by page charge payment. Therefore, and solely to indicate this fact, this article is hereby marked “advertisement” in accordance with 18 U.S.C. section 1734.

1
Hemler
ME
VLA proteins in the integrin family: structures, functions, and their role on leukocytes.
Annu Rev Immunol.
8
1990
365
400
2
Osborn
L
Hession
C
Tizard
R
et al
Direct expression cloning of vascular cell adhesion molecule 1, a cytokine-induced endothelial protein that binds to lymphocytes.
Cell.
59
1989
1203
1211
3
Salomon
D
Crisa
L
Mojcik
CF
Ishii
JK
Klier
FG
Shevach
EM
Vascular cell adhesion molecule-1 is expressed by cortical thymic epithelial cells and mediates thymocyte adhesion: implications for function VLA4 integrin in T-cell development.
Blood.
89
1997
2461
2471
4
Alon
R
Kassner
PD
Carr
MW
Finger
EB
Hemler
M
Springer
TA
The integrin VLA-4 supports tethering and rolling in flow on VCAM-1.
J Cell Biol.
127
1995
1485
1495
5
Arroyo
AG
Yang
JT
Rayburn
H
Hynes
RO
Differential requirements for α4 integrins in hematopoiesis.
Cell.
85
1996
997
1008
6
Pigott
R
Dillon
LP
Hemingway
I
Gearing
AJH
Soluble forms of E-selectin, ICAM-1, and VCAM-1 are present in the supernatant of cytokine activated culture endothelial cells.
Biochem Biophys Res Comm.
187
1992
584
589
7
Wellicome
SM
Kapahi
P
Mason
JC
Lebranchu
Y
Yarwood
H
Haskard
DO
Detection of a circulating form of vascular cell adhesion molecule-1: raised levels in rheumatoid arthritis and systemic lupus erythematosus.
Clin Exp Immunol.
92
1993
412
418
8
Leca
G
Mansur
SE
Bensussan
A
Expression of VCAM-1 (CD106) by a subset of TCR-gd-bearing lymphocyte clones: involvement of a metalloprotease in the specific hydrolytic release of the soluble isoform.
J Immunol.
154
1995
1069
1077
9
Kitani
A
Nakashima
N
Izumihara
T
et al
Soluble VCAM-1 induces chemotaxis of jurkat and synovial fluid cells bearing high affinity very late antigen-4.
J Immunol.
161
1998
4931
4938
10
Kitani
A
Nakashima
N
Matsuda
T
et al
T cells bound by vascular cell adhesion molecule-1/CD106 in synovial fluid in rheumatoid arthritis: inhibitory role of soluble vascular cell adhesion molecule-1 in T-cell activation.
J Immunol.
156
1996
2300
2308
11
Hughes
PE
Pfaff
M
Integrin affinity modulation.
Trends Cell Biol.
8
1998
359
364
12
Schwartz
MA
Schaller
MD
Ginsberg
MH
Integrins: emerging paradigms of signal transduction.
Ann Rev Cell Dev Biol.
11
1995
549
599
13
Bazzoni
G
Hemler
ME
Are changes in integrin affinity and conformation overemphasized?
Trends Biochem Sci.
23
1998
30
34
14
Kucik
DF
Dustin
ML
Miller
JM
Brown
EJ
Adhesion-activating phorbol esters increase the mobility of leukocyte integrin LFA-1 in cultured lymphocytes.
J Clin Invest.
97
1996
2139
2144
15
Stewart
M
Hogg
N
Regulation of leukocyte integrin function: affinity vs. avidity.
J Cell Biochem.
61
1996
554
561
16
Altieri
DC
Bader
R
Mannucci
PM
Edgington
TS
Oligospecificity of the cellular adhesion receptor MAC-1 encompasses an inducible recognition specificity for fibrinogen.
J Cell Biol.
107
1988
1893
1900
17
Faull
RJ
Kovach
NL
Harlan
JM
Ginsberg
MH
Stimulation of integrin-mediated adhesion of T lymphocytes and monocytes: two mechanisms with divergent biological consequences.
J Exp Med.
179
1994
1307
1316
18
Jakubowski
A
Rosa
MD
Bixler
S
Lobb
R
Burkly
LC
Vascular cell adhesion molecule (VCAM)-Ig fusion protein defines distinct affinity states of the very late antigen-4 (VLA-4) receptor.
Cell Adhes Commun.
3
1995
131
142
19
Giblin
PA
Hwang
ST
Katsumoto
TR
Rosen
SD
Ligation of L-selectin on T lymphocytes activates beta1 integrins and promotes adhesion to fibronectin.
J Immunol.
159
1997
3498
3507
20
Yednock
TA
Cannon
C
Vandevert
C
et al
α4β1 integrin-dependent cell adhesion is regulated by a low affinity receptor pool that is conformationally responsive to ligand.
J Biol Chem.
270
1995
28,740
28,750
21
Puzon-McLaughlin
W
Yednock
TA
Takada
Y
Regulation of conformation and ligand binding function of alpha5beta1 by the beta1 cytoplasmic domain.
J Biol Chem.
271
1996
16,580
16,585
22
Kamata
T
Puzon
W
Takada
Y
Identification of putative ligand-binding sites of the integrin α4β1 (VLA-4, CD49d/CD29).
Biochem J.
305
1995
945
951
23
LaFlamme
SE
Akiyama
SK
Yamada
KM
Regulation of fibronectin receptor distribution.
J Cell Biol.
117
1992
437
447
24
Bruder JT, Heidecker G, Rapp UR. Serum-, TPA-, and ras-induced expression from AP-1/Ets-driven promoters requires Raf-1 kinase. Genes Dev. 1992;545-556.
25
Hughes
PE
Renshaw
MW
Pfaff
M
et al
Suppression of integrin activation: a novel function of a Ras/Raf-initiated MAP kinase pathway.
Cell.
88
1997
521
530
26
Schoen
RC
Bentley
KL
Klebe
RJ
Monoclonal antibody against human fibronectin which inhibits cell attachment.
Hybridoma.
1
1982
99
108
27
Bowditch
RD
Halloran
CE
Aota
S
et al
Integrin αIIbβ3 (platelet GPIIb-IIIa) recognizes multiple sites in fibronectin.
J Biol Chem.
266
1991
23,323
23,328
28
Sanchez-Aparicio
P
Dominguez-Jimenez
C
Garcia-Pardo
A
Activation of the α4β1 integrin through the b1 subunit induces recognition of the RGDS sequence in fibronectin.
J Cell Biol.
126
1994
271
279
29
Fenczik
CA
Sethi
T
Ramos
JW
Hughes
PE
Ginsberg
MH
Complementation of dominant suppression implicates CD98 in integrin activation.
Nature.
370
1997
81
5
30
Osborn
L
Vassallo
C
Browning
BG
et al
Arrangement of domains, and amino acid residues required for binding of vascular cell adhesion molecule-1 to its counter-receptor VLA-4(α4β1).
J Cell Biol.
124
1994
601
608
31
Hession
C
Tizard
R
Vassallo
C
et al
Cloning of an alternate form of vascular cell adhesion molecule-1 (VCAM1).
J Biol Chem.
266
1991
6682
6685
32
Kassner
PD
Alon
R
Springer
TA
Hemler
M
Specialized functional properties of the integrin alpha4 cytoplasmic domain.
Mol Biol Cell.
6
1995
661
674
33
Weitzman
JB
Pujades
C
Hemler
ME
Integrin alpha chain cytoplasmic tails regulate “antibody-redirected” cell adhesion, independently of ligand binding.
Eur J Immunol.
27
1997
78
84
34
Yauch
RL
Felsenfeld
DP
Kraeft
SK
Chen
LB
Sheetz
MP
Hemler
ME
Mutational evidence for control of cell adhesion through integrin diffusion/clustering, independent of ligand binding.
J Exp Med
186
1997
1347
1355
35
Faull
RJ
Kovach
NL
Harlan
JM
Ginsberg
MH
Affinity modulation of integrin α5β1: regulation of the functional response by soluble fibronectin.
J Cell Biol.
121
1993
155
162
36
Trinchieri
G
Biology of natural killer cells.
Adv Immunol.
47
1990
187
376
37
Timonen
T
Natural killer cells: endothelial interactions, migration, and target cell recognition.
J Leukoc Biol.
62
1997
693
701
38
Palecek
SP
Loftus
JC
Ginsberg
MH
Horwitz
AF
Lauffenburger
DA
Integrin-ligand binding properties govern cell migration speed through cell-substratum adhesiveness.
Nature.
385
1997
537
540
39
Springer
TA
Traffic signals for lymphocyte recirculation and leukocyte emigration: the multistep paradigm.
Cell.
76
1994
301
314
40
Bradley
LM
Watson
SR
Lymphocyte migration into tissue: the paradigm derived from CD4 subsets.
Curr Opin Immunol.
8
1996
312
320
41
Salomon
DR
Mojcik
CF
Chang
AC
et al
Constitutive activation of integrin α4β1 defines a unique stage of human thymocyte development.
J Exp Med.
179
1994
1573
1584
42
Kerr
JFR
Wyllie
AH
Currie
AR
Apoptosis: a basic biological phenomenon with wide-ranging implications in tissue kinetics.
Br J Cancer.
26
1972
239
257
43
Wyllie
AH
Kerr
JFR
Currie
AR
Cell death: the significance of apoptosis.
Int Rev Cytol.
68
1980
251
306
44
Martin
SJ
O'Brien
GA
Nishioka
WK
et al
Proteolysis of fodrin (non-erythroid) during apoptosis.
J Biol Chem.
270
1995
6425
6428
45
Brancolini
C
Lazarevic
D
Rodriquez
J
Schneidner
C
Dismantling cell-cell contact during apoptosis is coupled to caspase-dependent proteolytic cleavage of catenin.
J Cell Biol.
139
1997
759
771
46
Wen
L-P
Fahrni
JA
Troie
S
Guan
J-L
Orth
K
Rosen
GD
Cleavage of focal adhesion kinase by caspases during apoptosis.
J Biol Chem.
272
1997
26,056
26,061
47
Widmann
C
Gibson
S
Johnson
G
Caspase-dependent cleavage of signaling proteins during apoptosis: a turn-off mechanism for anti-apoptotic signals.
J Biol Chem.
273
1998
7141
7147
48
Meredith
JE
Jr
Mu
Z
Saido
T
Du
X
Cleavage of the cytoplasmic domain of the integrin beta3 subunit during endothelial cell apoptosis.
J Biol Chem.
273
1998
19,525
19,531
49
Shattil
SJ
O'Toole
TE
Eigenthaler
M
et al
β3 endonexin, a novel polypeptide that interacts specifically with the cytoplasmic tail of the integrin β3 subunit.
J Cell Biol.
131
1995
807
816
Sign in via your Institution