Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

The New England Journal of Medicine

Review Articles

Mechanisms of Disease
F R A N K L I N H . E P S T E I N , M . D. , Editor

C HEMOKINES C HEMOTACTIC C YTOKINES T HAT M EDIATE I NFLAMMATION


ANDREW D. LUSTER, M.D., PH.D.

HE attraction of leukocytes to tissues is essential for inflammation and the host response to infection. The process is controlled by chemokines, which are chemotactic cytokines. This review introduces the burgeoning family of cytokines, with special emphasis on their role in the pathophysiology of disease and their potential as targets for therapy.
STRUCTURE AND FUNCTION OF CHEMOKINES

a- and b-chemokines, which contain four cysteines, appear to be the largest families. In the a-chemokines, one amino acid separates the first two cysteine residues (cysteineX amino acidcysteine, or CXC), whereas in the b-chemokines, the first two cysteine residues are adjacent to each other (cysteinecysteine, or CC) (Fig. 1). Two chemokines that do not fit into this classification, lymphotactin,9 with only two cysteines, and fractalkine,10 a membrane-bound glycoprotein in which the first two cysteine residues are separated by three amino acids (CXXXC) and the chemokine domain sits on a mucin-like stalk, may represent additional families. The a-chemokines can be further subdivided into those that contain the sequence glutamic acidleucinearginine near the N terminal (preceding the CXC sequence) and those that do not (Fig. 1).11 The a-chemokines containing the sequence are chemotactic for neutrophils, whereas those not containing the sequence act on lymphocytes. For example, IP-10 and MIG (monokine induced by interferon-g) attract activated T cells,12 and stromal-cellderived factor 1 acts on resting lymphocytes.13
Figure 1. Chemokines and Their Receptors. The chemokines are homologous 8-to-10-kd proteins that are subdivided into families on the basis of the relative position of the cysteine residues in the mature protein. In the a-chemokines, the first two cysteine residues are separated by a single amino acid (CXC), whereas in the b-chemokines, the first two cysteine residues are adjacent to each other (CC). The a-chemokines that contain the sequence glutamic acidleucinearginine preceding the CXC sequence are chemotactic for neutrophils, and those that do not contain this sequence act on lymphocytes. The C chemokine lymphotactin has only two cysteines in the mature protein, and the CXXXC chemokine fractalkine has three amino acids separating the first two cysteines. Chemokine receptors are G proteincoupled proteins that are expressed on subgroups of leukocytes. Four human CXC chemokine receptors (CXCR1 through CXCR4), eight human CC chemokine receptors (CCR1 through CCR8), and one human CXXXC chemokine receptor (CX3CR1) have been identified. MCP denotes monocyte chemoattractant protein, MIP macrophage inflammatory protein, RANTES regulated upon activation normal T-cell expressed and secreted, MDC macrophage-derived chemokine, HCC-1 hemofiltrate CC chemokine, TECK thymus-expressed chemokine, SDF-1 stromal-cellderived factor 1, TARC thymus and activation-regulated chemokine, ELC EBI1 (EpsteinBarr virusinduced gene 1)ligand chemokine, PARC pulmonary and activation-regulated chemokine, SLC secondary lymphoid-tissue chemokine, 6Ckine 6-cysteine chemokine, IP-10 interferon-inducible protein 10, MIG monokine induced by interferon-g, I-TAC interferon-inducible T-cell alpha chemoattractant, DC-CK1 dendritic-cell chemokine 1, LARC liver and activation-regulated chemokine, GCP granulocyte chemotactic protein, GRO growth-regulated oncogene, ENA-78 epithelial-cellderived neutrophil-activating peptide 78, NAP-2 neutrophil-activating peptide 2, and LIX lipopolysaccharide-induced CXC chemokine.

Over 40 chemokines have been identified to date, most of them in the past few years. The relations among chemokines were not initially appreciated, which led to an idiosyncratic nomenclature consisting of many acronyms. When initially identified, these proteins had no known biologic activity but were associated with inflammatory disease (e.g., platelet factor 41 and IP-10 [interferon-inducible protein of 10 kd]2). Only after interleukin-8,3 monocyte chemoattractant protein 1,4,5 and macrophage inflammatory protein 1a and 1b6 had been identified as leukocyte chemotactic factors was it clear that these proteins have in common important structural features and the ability to attract leukocytes. Chemokines are 8-to-10-kd proteins with 20 to 70 percent homology in amino acid sequences. They have been subdivided into families on the basis of the relative position of their cysteine residues.7,8 There are at least four families of chemokines, but only two have been extensively characterized. The

From the Infectious Disease Unit, Partners AIDS Research Center, Massachusetts General Hospital, and the Department of Medicine, Harvard Medical School both in Boston. Address reprint requests to Dr. Luster at the Infectious Disease Unit, Massachusetts General Hospital East, Bldg. 149, 13th St., Charlestown, MA 02129. 1998, Massachusetts Medical Society.

436

Fe b r u a r y 1 2 , 1 9 9 8 Downloaded from www.nejm.org at MCDANIEL COLLEGE on May 07, 2004. Copyright 1998 Massachusetts Medical Society. All rights reserved.

M E C H A N I S M S O F D I S E AS E

Chemokine
Chemokine receptor

Receptor

Cell Type

MCP-3, -4; MIP-1a; RANTES MCP-3, -4; eotaxin-1, -2; RANTES

CCR1 CCR3

Eosinophil

MCP-1, -2, -3, -4, -5 MCP-3, -4; eotaxin-1, -2; RANTES MCP-3, -4; MIP-1a; RANTES MCP-1, -2, -3, -4, -5 MIP-1a, MIP-1b, RANTES I-309 MDC, HCC-1, TECK Fractalkine SDF-1 MCP-3, -4; MIP-1a; RANTES MCP-1, -2, -3, -4, -5 TARC MIP-1a, MIP-1b, RANTES MIP-3b (ELC) PARC, SLC, 6CKine (Exodus-2) Fractalkine IP-10, MIG, I-TAC

CCR2 CCR3 CCR1 CCR2 CCR5 CCR8 ? CX3CR1 CXCR4 CCR1 CCR2 CCR4 CCR5 CCR7 ? CX3CR1 CXCR3 ? ? CXCR4 CCR1 CCR2 CCR3 CCR4 CCR5 CCR6 ? CXCR4 CXCR1 CXCR2

Basophil

CC

C CC C

Monocyte

Activated T cell

C C

PARC, DC-CK1 Lymphotactin SDF-1

Resting T cell

CXC

C CXC C

MCP-3, -4; MIP-1a; RANTES MCP-1, -2, -3, -4, -5 MCP-3, -4; eotaxin-1, -2; RANTES TARC MIP-1a, MIP-1b, RANTES MIP-3a (LARC, Exodus-1) MDC, TECK SDF-1

Dendritic cell

Glutamic acid leucine arginine

Interleukin-8, GCP-2 Interleukin-8, GCP-2; GRO-a, -b, -g; ENA-78; NAP-2; LIX MCP-1, -2, -3, -4, -5 MIP-1a, MIP-1b, RANTES

Neutrophil

CCR2 CCR5

CXXXC
Chemokine domain

Mucin-like domain

Cytoplasmic domain

Natural killer cell CX3CR1 CXCR3

CXXXC C
Fractalkine IP-10, MIG, I-TAC

Vo l u m e 3 3 8 Downloaded from www.nejm.org at MCDANIEL COLLEGE on May 07, 2004. Copyright 1998 Massachusetts Medical Society. All rights reserved.

Nu m b e r 7

437

The New England Journal of Medicine

The b-chemokines, in general, do not act on neutrophils but attract monocytes, eosinophils, basophils, and lymphocytes with variable selectivity. Structurally, the b-chemokines can be subdivided into two families (Fig. 1): the monocyte-chemoattractant-proteineotaxin family, containing the five monocyte chemoattractant proteins and eotaxin, which are approximately 65 percent identical to each other, and all other b-chemokines.14 As with the CXC family, the N-terminal amino acids preceding the CC residues of b-chemokines are critical components of the biologic activity and leukocyte selectivity of these chemokines. For example, the addition of a single amino-acid residue at the amino terminal of monocyte chemoattractant protein 1 reduces its biologic activity on monocytes by 100 to 1000 times,15 and the deletion of a single amino acid from that region converts the chemokine from an activator of basophils to an eosinophil chemoattractant.16 Several chemokines undergo N-terminal proteolytic processing after secretion, which alters their activity. For example, the inactive platelet granule chemokine, platelet basic protein, is processed at the N terminal by monocyte proteases to form neutrophil chemoattractant neutrophil-activating peptide 2.17 This may reflect a general mechanism that allows local factors to regulate and amplify chemokine activity.
CHEMOKINE RECEPTORS

Chemokines induce cell migration and activation by binding to specific G-proteincoupled cell-surface receptors on target cells.18,19 Four human CXC chemokine receptors (CXCR1 through CXCR4), eight human CC chemokine receptors (CCR1 through CCR8), and one human CXXXC chemokine receptor (CX3CR1) have been identified (Fig. 1). Chemokine receptors are expressed on different types of leukocytes. Some receptors are restricted to certain cells (e.g., CXCR1 is predominantly restricted to neutrophils), whereas others are more widely expressed (e.g., CCR2 is expressed on monocytes, T cells, natural killer cells, dendritic cells, and basophils). In addition, chemokine receptors are constitutively expressed on some cells, whereas they are inducible on others. CCR1 and CCR2 are constitutively expressed on monocytes but are expressed on lymphocytes only after stimulation by interleukin-2.20 In addition, some constitutive chemokine receptors can be down-regulated; CCR2 is down-regulated by lipopolysaccharide, making the cells unresponsive to monocyte chemoattractant protein 1 (which activates only this receptor), but it remains responsive to macrophage inflammatory protein 1a (which activates CCR1 and CCR5).21 In contrast, the expression of other chemokine receptors is restricted to a cell state of activation and differentiation. For example, CXCR3 is expressed on
438
Fe b r u a r y 1 2 , 1 9 9 8

activated T lymphocytes of the T helper type 1 (Th1) phenotype, whereas CCR3, in addition to being expressed on eosinophils and basophils, is preferentially expressed on activated lymphocytes of the T helper type 2 (Th2) phenotype.22 In this way, transient up-regulation of chemokine receptors on leukocytes allows for the selective amplification of either a cell-mediated Th1-type immune response or an allergic Th2-type response. Some chemokine receptors are also expressed on nonhematopoietic cells, including neurons, astrocytes, epithelial cells, and endothelial cells. This suggests that the chemokine system has other roles in addition to leukocyte chemotaxis. Although most chemokine receptors bind more than one chemokine, CC receptors bind only CC chemokines and CXC receptors bind only CXC chemokines. This ligandreceptor restriction may be related to structural differences between CC and CXC chemokines, which have similar primary, secondary, and tertiary structures but different quaternary structures.23 Chemokine receptors, like other members of the family of G-proteincoupled receptors, are functionally linked to phospholipases through G proteins.18,19,24 Many chemokine-induced signaling events are inhibited by Bordetella pertussis toxin, suggesting that chemokine receptors are linked to G proteins of the Gi class. Receptor activation leads to a cascade of cellular activation, including the generation of inositol triphosphate, the release of intracellular calcium, and the activation of protein kinase C.23 Chemokine-receptor signaling also activates small guanosine triphosphatebinding proteins of the Ras and Rho families.25 Rho proteins are involved in cell motility through regulation of actin-dependent processes such as membrane ruffling, pseudopod formation, and assembly of focal adhesion complexes. Thus, chemokine receptors activate multiple intracellular signaling pathways that regulate the intracellular machinery necessary to propel the cell in its chosen direction. Chemokines also interact with two types of nonsignaling molecules. One is the erythrocyte chemokine receptor, called DARC (Duffy antigen receptor for chemokines).26 This receptor, known since the 1950s as the determinant of the Duffy blood group, is expressed on erythrocytes and endothelial cells. Although DARC is structurally related to chemokine receptors, it is distinctive in that both CXC and CC chemokines bind to it and chemokine binding does not induce calcium flux. This receptor may function as a sink for chemokines, clearing them from the circulation. The second type is a group of heparan sulfate proteoglycans. Chemokines are basic proteins, and they bind avidly to negatively charged heparin and heparan sulfate.27,28 Heparan sulfate proteoglycans capture chemokines in the extracellu-

Downloaded from www.nejm.org at MCDANIEL COLLEGE on May 07, 2004. Copyright 1998 Massachusetts Medical Society. All rights reserved.

M E C H A N I S M S O F D I S E AS E

lar matrix and on the surface of endothelial cells, a process that may serve to establish a local concentration gradient from the point source of chemokine secretion.29
ROLE IN LEUKOCYTE MOVEMENT

Chemokines are thought to provide the directional cues for the movement of leukocytes in development, homeostasis, and inflammation. Leukocyte extravasation from the blood into the tissues is a regulated multistep process involving a series of coordinated interactions between leukocytes and endothelial cells (Fig. 2).30,31 Several families of molecular regulators, such as selectins, integrins, and chemokines, are thought to control different aspects of this process. Selectins facilitate the movement of leukocytes along the surface of endothelial cells (rolling). Chemokines are thought to provide the signals that convert the low-affinity, selectin-mediated interaction into the higher-affinity, integrinmediated interaction that leads to extravasation of leukocytes. Chemokines are believed to control the homeostatic circulation of leukocytes through tissues. The continuous recirculation of lymphocytes through the blood, tissues, and lymphatics in an organized manner brings naive lymphocytes into the lymph nodes, where they encounter antigen and are transformed in memory lymphocytes that migrate into inflamed tissue to ensure immunity. Macrophages, eosinophils, and mast cells also migrate into tissue. Although these cells are produced in the bone marrow, they reside primarily in other tissues. The role of chemokines in regulating the movement of cells into tissues has begun to be elucidated on the basis of studies in mice deficient in a particular chemokine. For example, stromal-cellderived factor 1 is critical for the migration of myeloid precursors from the fetal liver to the bone marrow,32 and eotaxin is important in the recruitment of eosinophils into tissues.33 The dramatic increase in the secretion of chemokines during inflammation results in the selective recruitment of leukocytes into inflamed tissue. Chemokines have been detected during inflammation in most organs, including the skin, brain, joints, meninges, lungs, blood vessels, kidneys, and gastrointestinal tract. They have also been identified in many types of cells during inflammation in these organs, suggesting that most, if not all, cells can secrete chemokines, given the appropriate stimulus. The main stimuli for chemokine production are early proinflammatory cytokines, such as interleukin-1 and tumor necrosis factor a, bacterial products, such as lipopolysaccharide, and viral infection.7 In addition, interferon-g and interleukin-4, products of Th1 and Th2 lymphocytes, respectively, can induce the production of chemokines and also synergize with in-

terleukin-1 and tumor necrosis factor a to stimulate chemokine secretion.34,35 If proinflammatory cytokines stimulate secretion of many of the same chemokines, how can an inflammatory response be specific? In asthma, for example, eosinophils accumulate in the airways, whereas the host response to bacterial pneumonia is dominated by neutrophils. As pointed out above, the recruitment of leukocytes into tissues is a multistep process in which chemokines participate but do not act alone. Chemokines often act in concert with other cytokines to cause tissue infiltration, by increasing the circulating pool of a given leukocyte and up-regulating particular adhesion molecules, as well as increasing leukocyte responsiveness to a chemokine. For example, eotaxin and interleukin-5 together cause tissue eosinophilia,36 and interleukin-5 and interleukin-3 prime basophils to release histamine and leukotriene after stimulation by monocyte chemoattractant protein.37
ROLE IN INFLAMMATORY DISEASES

The secretion of chemokines has been detected in a wide variety of diseases (Fig. 3).7 It is likely that in these diseases chemokines cause the accumulation and activation of leukocytes in tissues. The capacity to control precisely the movement of inflammatory cells suggests that the various chemokines and their receptors might provide novel targets for therapeutic interventions. The type of inflammatory infiltrate that characterizes a specific disease is controlled, in part, by the subgroup of chemokines expressed in the diseased tissue (Fig. 3). For example, in many acute disease processes, such as bacterial pneumonia and the acute respiratory distress syndrome, there is a massive influx of neutrophils into the tissue. The concentration of potent neutrophil chemoattractants, such as interleukin-8, is increased in bronchoalveolar fluid from patients with these pulmonary diseases.38 In viral meningitis monocytes and lymphocytes are recruited into the tissue, concentrations of chemokines that are active on these cells, such as IP-10 and monocyte chemoattractant protein 1, are increased in cerebrospinal fluid, and the increases are correlated with the extent of mononuclear-cell infiltration of the meninges.39 Tissue infiltration by lymphocytes and macrophages occurs in many chronic diseases. Activated lymphocytes accumulate in the granulomatous lesions that are characteristic of tuberculoid leprosy and sarcoidosis, and high concentrations of IP-10 have been detected.40 Moreover, the concentration of IP-10 in bronchoalveolar fluid from patients with active sarcoidosis is correlated with the number of activated T lymphocytes in the fluid (unpublished observations). In atherosclerosis, macrophages and lymphocytes are the chief inflammatory cells found
Vo l u m e 3 3 8 Nu m b e r 7

439

Downloaded from www.nejm.org at MCDANIEL COLLEGE on May 07, 2004. Copyright 1998 Massachusetts Medical Society. All rights reserved.

The New England Journal of Medicine

Red cells

Lumen

DARC

Integrin

Adherence
Activated integrin

Endothelium

Activation

Rolling Extravasation
Intercellular adhesion molecule Mucin Cell-surface heparan sulfate proteoglycans Selectin

Inflammatory trigger (e.g., infection, allergen, autoantigen, alloantigen, tumor) Activated recruited leukocyte

Chemokine receptor

Interleukin-1 receptor or tumor necrosis factor receptor

Matrix heparan sulfate proteoglycans Chemokine

Leukocyte

Interleukin-1 and tumor necrosis factor

Figure 2. Chemokine Regulation of Leukocyte Movement. Chemokines are secreted at sites of inflammation and infection by resident tissue cells, resident and recruited leukocytes, and cytokine-activated endothelial cells. Chemokines are locally retained on matrix and cell-surface heparan sulfate proteoglycans, establishing a chemokine concentration gradient surrounding the inflammatory stimulus, as well as on the surface of the overlying endothelium. Leukocytes rolling on the endothelium in a selectin-mediated process are brought into contact with chemokines retained on cell-surface heparan sulfate proteoglycans. Chemokine signaling activates leukocyte integrins, leading to firm adherence and extravasation. The recruited leukocytes are activated by local proinflammatory cytokines and may become desensitized to further chemokine signaling because of high local concentrations of chemokines. The Duffy antigen receptor for chemokines (DARC), a nonsignaling erythrocyte chemokine receptor, functions as a sink, removing chemokines from the circulation and thus helping to maintain a tissuebloodstream chemokine gradient.

440

Fe b r u a r y 1 2 , 1 9 9 8 Downloaded from www.nejm.org at MCDANIEL COLLEGE on May 07, 2004. Copyright 1998 Massachusetts Medical Society. All rights reserved.

M E C H A N I S M S O F D I S E AS E

Inflammatory Disease
Acute respiratory distress syndrome

Infiltrate
Neutrophil

Chemokine
Interleukin-8; GRO-a, -b, -g; ENA-78 MCP-1, -4; MIP-1a; eotaxin; RANTES

Figure 3. Role of Chemokines in Various Inflammatory Diseases. Inflammatory diseases are characterized by the selective accumulation of leukocyte subgroups, a process controlled by the expression of certain chemokines. Each disease has a characteristic inflammatory infiltrate in which chemokine messenger RNA or protein concentrations have been shown to be up-regulated. The chemokine abbreviations are explained in the legend to Figure 1.

Asthma

Eosinophil, T cell, monocyte, basophil

Bacterial pneumonia

Neutrophil

Interleukin-8, ENA-78

Sarcoidosis

T cell, monocyte

IP-10

Glomerulonephritis

Monocyte, T cell, neutrophil

MCP-1, RANTES, IP-10

Rheumatoid arthritis Monocyte, neutrophil Osteoarthritis

MIP-1a, MCP-1, interleukin-8, ENA-78 MIP-1b

T cell, monocyte Atherosclerosis Monocyte, neutrophil, T cell, eosinophil

MCP-1, -4; IP-10

Inflammatory bowel disease

MCP-1, MIP-1a, eotaxin, IP-10, interleukin-8

T cell, neutrophil

Psoriasis Neutrophil, monocyte

MCP-1, IP-10, MIG, GRO-b, interleukin-8 Interleukin-8; GRO-a; MCP-1; MIP-1a, -1b MCP-1, IP-10

Bacterial meningitis

Viral meningitis

T cell, monocyte

in the diseased blood vessels. As progenitors of lipidladen foam cells and a source of growth factors that mediate intimal hyperplasia, these inflammatory cells may be central to the pathogenesis of atherosclerosis. Although the mechanism of monocyte recruitment in atherosclerotic lesions is unknown, monocyte chemoattractant protein 1 has been detected in diseased carotid arteries but not in normal carotid arteries.41 In asthma, rhinitis, and atopic dermatitis, there is selective accumulation and activation of eosinophils and mast cells, and mediators derived from these cells participate in the pathogenesis of these allergic diseases.42 Agents that induce the release of histamine from mast cells and basophils, so-called histamine-releasing factors, play an important part. Chemokines, especially eotaxin and the monocyte chemoattractant proteins, are potent eosinophil chemoattractants and histamine-releasing factors, making them particularly important in allergic inflammation.14 In fact, these chemokines may be the main histamine-releasing factors in the absence of antigen and IgE antibody. Many chemokines have been detected in the airways of patients with asthma.43-45 In addition, several chemokines that act on eosinophils are increased in the epithelial tissue in patients with atopic dermatitis, allergic rhinitis, or asthma after an antigen challenge, making it likely that these chemokines are a molecular link between antigen-specific immune activation and the migration of eosinophils into tissues.34,46,47 Ulcerative colitis and Crohns disease are characterized by chronic inflammation with superimposed acute inflammatory exacerbations. In the chronic phase, macrophages and lymphocytes infiltrate the bowel, and in the acute phase, neutrophils and perhaps eosinophils leave the circulation and enter the intestinal mucosa. Many chemokines are markedly increased in intestinal tissue from patients with ulcerative colitis or Crohns disease.35,48,49 In psoriasis, the lesions contain neutrophils and activated T cells, and the neutrophil chemoattractants interleukin-8 and GRO-a (growth-related oncogene a, although it is not an oncogene). The activated T-cell chemoattractants IP-10 and monocyte chemoattractant protein 1 are present in psoriatic plaques but not in normal skin.40,50 In addition, successful treatment of

Vo l u m e 3 3 8 Downloaded from www.nejm.org at MCDANIEL COLLEGE on May 07, 2004. Copyright 1998 Massachusetts Medical Society. All rights reserved.

Nu m b e r 7

441

The New England Journal of Medicine

psoriatic plaques results in decreased IP-10 in diseased skin.40


ROLE IN INFECTIOUS DISEASES

another intriguing link between the chemokine system and diseases in humans.
MODULATION OF ANGIOGENESIS, TUMOR GROWTH, AND STEM-CELL PROLIFERATION

Pathogenic organisms bind to receptors on leukocytes to gain access to the cytoplasm and nucleus of cells (e.g., EpsteinBarr virus binds to complement receptor 3, and rhinoviruses bind to intercellular adhesion molecule 1). The chemokine receptors serve as coreceptors for two important human pathogens, plasmodium and the human immunodeficiency virus (HIV). Plasmodium vivax binds to the DARC receptor on erythrocytes,26 and HIV binds to several chemokine receptors. By facilitating entry into cells, these receptors determine viral tropism.51-57 CXCR4 is a coreceptor for strains of HIV type 1 (HIV-1) that infect T-cell lines (T-tropic strains), and CCR5 is a coreceptor for HIV-1 isolates that infect macrophages and activated T cells (M-tropic strains) (Fig. 4). RANTES (regulated upon activation normal T-cell expressed and secreted) and macrophage inflammatory proteins 1a and 1b, which are CCR5 ligands, and stromal-cellderived factor 1, a CXCR4 ligand, block the entry of M-tropic and T-tropic HIV, respectively, into cells.58-60 The importance of chemokine receptors in the pathophysiology of HIV infection became apparent when it was discovered that a polymorphic variant of CCR5 could explain the observation that certain persons who are at high risk for HIV-1 infection remain uninfected.61-63 In persons who are homozygous for a 32-base-pair deletion in the gene for CCR5, a functional CCR5 protein cannot be synthesized, and such persons are not found in HIV-1 positive cohorts. In addition, cells from persons with this mutation cannot be infected with HIV-1 in vitro.64 Furthermore, in persons who are heterozygous for the mutation, the rate of progression of HIV-1 infection is slower than in those without the mutation.61 Although the molecular mechanism of the fusion event is not known, HIV glycoprotein 120, CD4, and a chemokine receptor associate on the cell membrane before HIV enters the cells.65-67 Pathogenic organisms express cytokines and cytokine receptors that are important in infection (e.g., poxviruses encode functional interleukin-1b and interferon-g receptors). Many of the herpesviruses express chemokine-receptor homologues, and many of these homologues can bind chemokines. Kaposis sarcomaassociated herpesvirus 8 was recently shown to encode a constitutively (agonist-independent) active chemokine receptor that stimulated cellular proliferation.68 In addition, human herpesvirus 6, Kaposis sarcomaassociated herpesvirus 8, and the molluscum contagiosum virus encode CC chemokine homologues. Although the role of these viral chemokine and chemokine-receptor homologues in the infectious process is unknown, they represent
442
Fe b r u a r y 1 2 , 1 9 9 8

Chemokines can also modulate angiogenesis and tumor growth and inhibit stem-cell proliferation. Platelet factor 469 and IP-1028,70 inhibit neovascularization, tumor growth, and metastasis.71 In contrast, interleukin-8 promotes angiogenesis and tumor metastasis.71,72 The mechanism underlying the inhibition of angiogenesis may be related to the ability of certain chemokines to displace other growth factors, such as basic fibroblast growth factor and transforming growth factor a, from heparan sulfate sites on endothelial cells. The ability of certain chemokines to inhibit stemcell proliferation may have therapeutic applications.73 Macrophage inflammatory protein 1a and analogues with antistem-cell but not proinflammatory activity are being developed as adjuncts to cancer chemotherapy. If they can selectively arrest stem-cell replication, then high-dose chemotherapy may be better tolerated because of decreased stem-cell toxicity. Monocyte chemoattractant proteins 1, 2, and 3 have been isolated from glioma and osteosarcoma cells.4,74 Many other chemokines are produced by tumor cells in vitro and are present in human tumors.34,35,71 However, the overall role of chemokines in tumor biology is unclear. Tumor-associated leukocytes may stimulate or inhibit tumor growth. They can promote growth by supplying growth factors and promoting angiogenesis and can inhibit growth by enhancing the host response to the tumor.
STUDIES IN ANIMALS

Animal models of disease have been useful in studying the relation between chemokine expression and pathophysiologic processes.75 In certain inflammatory responses, a single chemokine plays a major part in the recruitment of a subgroup of leukocytes. For example, there are numerous chemokines that act on neutrophils, but a neutralizing antibody against interleukin-8 completely blocks reperfusion-associated lung injury in rabbits.76 In addition, even though other b-chemokines have many of the actions of macrophage inflammatory protein 1a, including an ability to bind to the same receptor, in studies of mice with a deletion of the gene for this chemokine, myocarditis did not develop when the mice were infected with coxsackievirus B3, and there was a significant reduction in the histologic degree of pneumonitis after infection with influenzavirus.77 The recruitment of leukocytes in other inflammatory responses depends on multiple chemokines. In animals with allergic pulmonary inflammation, the expression of eotaxin, macrophage inflammatory

Downloaded from www.nejm.org at MCDANIEL COLLEGE on May 07, 2004. Copyright 1998 Massachusetts Medical Society. All rights reserved.

M E C H A N I S M S O F D I S E AS E

M-tropic HIV

M-tropic HIV

T-tropic HIV

T-tropic HIV

gp120 V3 loop V3 loop

MIP-1a, MIP-1b, or RANTES

gp120 V3 loop V3 loop SDF-1

CD4 Cell membrane

CCR5

CD4

CCR5

CD4

CXCR4

CD4

CXCR4

Entry of M-tropic HIV strains

Entry of M-tropic HIV strains blocked

Entry of T-tropic HIV strains

Entry of T-tropic HIV strains blocked

Figure 4. Chemokine Receptors as Obligate Coreceptors for HIV Entry into Cells and Chemokine Inhibition of HIV Entry. HIV glycoprotein 120 (gp120) binds to CD4, resulting in a conformational change that exposes the V3 loop in gp120 and permits subsequent interaction with a chemokine receptor. To gain entry into cells, macrophage-tropic (M-tropic) HIV-1 uses CCR5 predominantly, and the T-celltropic (T-tropic) HIV-1 uses CXCR4 predominantly. Macrophage inflammatory proteins (MIP) 1a and 1b and the RANTES (regulated upon activation normal T-cell expressed and secreted) chemokine, ligands for CCR5, block M-tropic HIV-1 from entering cells. Stromal-cellderived factor 1 (SDF-1), a ligand for CXCR4, blocks T-tropic HIV-1 from entering cells.

protein 1a, and monocyte chemoattractant proteins 1, 3, and 5 precedes the massive airway recruitment of mononuclear cells and eosinophils.14,78-80 On the basis of studies using antibodies that inhibit the action of these chemokines and studies of mice with a targeted disruption of the eotaxin gene, it is clear that all these chemokines participate in the recruitment of eosinophils into the airways.33,80-82 Other studies in animals suggest that although a given chemokine may be only partly responsible for the recruitment of a subgroup of leukocytes, it may be largely responsible for certain pathophysiologic aspects of a disease. For example, in animals with acute immune-complexmediated glomerulonephritis, the glomerular lesions contain neutrophils, and there is extensive fusion of the glomerular foot processes, with increased urinary protein excretion. The administration of an antiinterleukin-8 antibody reduces neutrophil infiltration by only 40 percent but completely prevents the fusion of epithelial foot processes and urinary protein excretion.83

Certain chemokines have a critical role in development. Mice deficient in stromal-cellderived factor 1 die perinatally, with defects in B-cell lymphopoiesis and the recruitment of hematopoietic progenitors from the fetal liver into the bone marrow, as well as a ventricular septal defect.32 This dramatic phenotype underscores the possibility that aside from directing the migration of leukocytes to sites of inflammation and infection, chemokines may play an important part in orchestrating the movement of cells during development.
CONCLUSIONS

Chemokines appear to have the capacity to control precisely the movement of leukocytes. The roles of chemokines in the pathophysiology of disease are still being defined, but there is growing evidence from studies in animals that the neutralization of chemokine activity may have therapeutic value. Chemokine or chemokine-receptor antagonists may inhibit autoimmune, allergic, and septic processes. In
Vo l u m e 3 3 8 Nu m b e r 7

443

Downloaded from www.nejm.org at MCDANIEL COLLEGE on May 07, 2004. Copyright 1998 Massachusetts Medical Society. All rights reserved.

The New England Journal of Medicine

addition, chemokines may augment the host response to infection, tumors, and vaccines. Chemokines in combination with other cytokines may provide more effective antitumor therapy than either alone, as shown in a study of lymphotactin combined with interleukin-2.84 Finally, chemokines or their analogues may be clinically useful as inhibitors of HIV-1 infection and disease progression. The chemokines are a fascinating family of cytokines that we are only beginning to understand. It is likely that additional chemokines and chemokine receptors will soon be discovered. The challenge for the future will be to understand the role of chemokines in the pathophysiology of disease.
Supported by grants from the National Institutes of Health (R01CA69212 and R01-AI40618). Dr. Luster is the recipient of a Cancer Research Institute Benjamin Jacobson Family Investigator Award and a Culpeper Medical Scholars Award.

REFERENCES
1. Deuel TF, Keim PS, Farmer M, Heinrikson RL. Amino acid sequence of human platelet factor 4. Proc Natl Acad Sci U S A 1977;74:2256-8. 2. Luster AD, Unkeless JC, Ravetch JV. Gamma-interferon transcriptionally regulates an early-response gene containing homology to platelet proteins. Nature 1985;315:672-6. 3. Yoshimura T, Matsushima K, Tanaka S, et al. Purification of a human monocyte-derived neutrophil chemotactic factor that has peptide sequence similarity to other host defense cytokines. Proc Natl Acad Sci U S A 1987; 84:9233-7. 4. Yoshimura T, Robinson EA, Tanaka S, Appella E, Kuratsu J, Leonard EJ. Purification and amino acid analysis of two human glioma-derived monocyte chemoattractants. J Exp Med 1989;169:1449-59. 5. Matsushima K, Larsen CG, DuBois GC, Oppenheim JJ. Purification and characterization of a novel monocyte chemotactic and activating factor produced by a human myelomonocytic cell line. J Exp Med 1989;169: 1485-90. 6. Wolpe SD, Davatelis G, Sherry B, et al. Macrophages secrete a novel heparin-binding protein with inflammatory and neutrophil chemokinetic properties. J Exp Med 1988;167:570-81. 7. Baggiolini M, Dewald B, Moser B. Interleukin-8 and related chemotactic cytokines CXC and CC chemokines. Adv Immunol 1994;55:97179. 8. Idem. Human chemokines: an update. Annu Rev Immunol 1997;15: 675-705. 9. Kelner GS, Kennedy J, Bacon KB, et al. Lymphotactin: a cytokine that represents a new class of chemokine. Science 1994;266:1395-9. 10. Bazan JF, Bacon KB, Hardiman G, et al. A new class of membranebound chemokine with a CX3C motif. Nature 1997;385:640-4. 11. Clark-Lewis I, Schumacher C, Baggiolini M, Moser B. Structure-activity relationships of interleukin-8 determined using chemically synthesized analogs: critical role of NH2-terminal residues and evidence for uncoupling of neutrophil chemotaxis, exocytosis, and receptor binding activities. J Biol Chem 1991;266:23128-34. 12. Loetscher M, Gerber B, Loetscher P , et al. Chemokine receptor specific for IP10 and mig: structure, function, and expression in activated T-lymphocytes. J Exp Med 1996;184:963-9. 13. Bleul CC, Fuhlbrigge C, Casasnovas JM, Aiuti A, Springer TA. A highly efficacious lymphocyte chemoattractant, stromal cell-derived factor 1 (SDF-1). J Exp Med 1996;184:1101-9. 14. Luster AD, Rothenberg ME. Role of the monocyte chemoattractant protein and eotaxin subfamily of chemokines in allergic inflammation. J Leukoc Biol 1997;62:620-33. 15. Gong J-H, Clark-Lewis I. Antagonists of monocyte chemoattractant protein 1 identified by modification of functionally critical NH2-terminal residues. J Exp Med 1995;181:631-40. 16. Weber M, Uguccioni M, Baggiolini M, Clark-Lewis I, Dahinden CA. Deletion of the NH2-terminal residue converts monocyte chemotactic protein 1 from an activator of basophil mediator release to an eosinophil chemoattractant. J Exp Med 1996;183:681-5. 17. Walz A, Dewald B, von Tscharner V, Baggiolini M. Effects of the neutrophil-activing peptide NAP-2, platelet basic protein, connective tissue-

activating peptide III and platelet factor 4 on human neutrophils. J Exp Med 1989;170:1745-50. 18. Premack BA, Schall TJ. Chemokine receptors: gateways to inflammation and infection. Nat Med 1996;2:1174-8. 19. Murphy PM. The molecular biology of leukocyte chemoattractant receptors. Annu Rev Immunol 1994;12:593-633. 20. Loetscher P, Seitz M, Baggiolini M, Moser B. Interleukin-2 regulates CC chemokine receptor expression and chemotactic responsiveness in T lymphocytes. J Exp Med 1996;184:569-77. 21. Sica A, Saccani A, Borsatti A, et al. Bacterial lipopolysaccharide rapidly inhibits expression of C-C chemokine receptors in human monocytes. J Exp Med 1997;185:969-74. 22. Sallusto F, Mackay CR, Lanzavecchia A. Selective expression of the eotaxin receptor CCR3 by human T helper 2 cells. Science 1997;277:2005-7. 23. Lodi PJ, Garrett DS, Kuszewski J, et al. High-resolution solution structure of the beta chemokine hMIP-1b by multidimensional NMR. Science 1994;263:1762-7. 24. Bokoch GM. Chemoattractant signaling and leukocyte activation. Blood 1995;86:1649-60. 25. Laudanna C, Campbell JJ, Butcher EC. Role of Rho in chemoattractant-activated leukocyte adhesion through integrins. Science 1996;271: 981-3. 26. Horuk R, Chitnis CE, Darbonne WC, et al. A receptor for the malarial parasite Plasmodium vivax: the erythrocyte chemokine receptor. Science 1993;261:1182-4. 27. Rot A. Endothelial cell binding of NAP-1/IL-8: role in neutrophil emigration. Immunol Today 1992;13:291-4. 28. Luster AD, Greenberg SM, Leder P . The IP-10 chemokine binds to a specific cell surface heparan sulfate site shared with platelet factor 4 and inhibits endothelial cell proliferation. J Exp Med 1995;182:219-31. 29. Tanaka Y, Adams DH, Hubscher S, Hirano H, Siebenlist U, Shaw S. T-cell adhesion induced by proteoglycan-immobilized cytokine MIP-1b. Nature 1993;361:79-82. 30. Butcher EC. Leukocyte-endothelial cell recognition: three (or more) steps to specificity and diversity. Cell 1991;67:1033-6. 31. Springer TA. Traffic signals for lymphocyte recirculation and leukocyte emigration: the multistep paradigm. Cell 1994;76:301-14. 32. Nagasawa T, Hirota S, Tachibana K, et al. Defects of B-cell lymphopoiesis and bone-marrow myelopoiesis in mice lacking the CXC chemokine PBSF/SDF-1. Nature 1996;382:635-8. 33. Rothenberg ME, MacLean JA, Pearlman E, Luster AD, Leder P . Targeted disruption of the chemokine eotaxin partially reduces antigeninduced tissue eosinophilia. J Exp Med 1997;185:785-90. 34. Garcia-Zepeda EA, Combadiere C, Rothenberg ME, et al. Human monocyte chemoattractant protein (MCP)-4 is a novel CC chemokine with activities on monocytes, eosinophils, and basophils induced in allergic and nonallergic inflammation that signals through the CC chemokine receptors (CCR)-2 and -3. J Immunol 1996;157:5613-26. 35. Garcia-Zepeda EA, Rothenberg ME, Ownbey RT, Celestin J, Leder P, Luster AD. Human eotaxin is a specific chemoattractant for eosinophil cells and provides a new mechanism to explain tissue eosinophilia. Nat Med 1996;4:449-56. 36. Collins PD, Marleau S, Griffiths-Johnson DA, Jose PJ, Williams TJ. Cooperation between interleukin-5 and the chemokine eotaxin to induce eosinophil accumulation in vivo. J Exp Med 1995;182:1169-74. 37. Bischoff SC, Brunner T, De Weck AL, Dahinden CA. Interleukin-5 modifies histamine release and leukotriene generation by human basophils in response to diverse agonists. J Exp Med 1990;172:1577-82. 38. Chollet-Martin S, Montravers P, Gibert C, et al. High levels of interleukin-8 in the blood and alveolar spaces of patients with pneumonia and adult respiratory distress syndrome. Infect Immun 1993;61:4553-9. 39. Lahrtz F, Piali L, Nadal D, et al. Chemokines in viral meningitis: chemotactic cerebrospinal fluid factors include MCP-1 and IP-10 for monocytes and activate T lymphocytes. Eur J Immunol 1997;27:2484-9. 40. Gottlieb AB, Luster AD, Posnett DN, Carter DM. Detection of a gamma interferon-induced protein IP-10 in psoriatic plaques. J Exp Med 1988;168:941-8. 41. Nelken NA, Coughlin SR, Gordon D, Wilcox JN. Monocyte chemoattractant protein-1 in human atheromatous plaques. J Clin Invest 1991; 88:1121-7. 42. Bousquet J, Chanez P, Lacoste JY, et al. Eosinophilic inflammation in asthma. N Engl J Med 1990;323:1033-9. 43. Sousa AR, Lane SJ, Nakhosteen JA, Yoshimura T, Lee TH, Poston RN. Increased expression of the monocyte chemoattractant protein-1 in bronchial tissue from asthmatic subjects. Am J Respir Cell Mol Biol 1994;10: 142-7. 44. Humbert M, Ying S, Corrigan C, et al. Bronchial mucosal expression of the genes encoding chemokines RANTES and MCP-3 in symptomatic atopic and nonatopic asthmatics: relationship to the eosinophil-active cy-

444

Fe b r u a r y 1 2 , 1 9 9 8 Downloaded from www.nejm.org at MCDANIEL COLLEGE on May 07, 2004. Copyright 1998 Massachusetts Medical Society. All rights reserved.

M E C H A N I S M S O F D I S E AS E

tokines interleukin (IL)-5, granulocyte macrophage-colony-stimulating factor, and IL-3. Am J Respir Cell Mol Biol 1997;16:1-8. 45. Lamkhioued B, Renzi PM, Abi-Younes S, et al. Increased expression of eotaxin in bronchoalveolar lavage and airways of asthmatics contributes to the chemotaxis of eosinophils to the site of inflammation. J Immunol 1997;159:4593-601. 46. Minshall EM, Cameron L, Lavigne F, et al. Eotaxin mRNA and protein expression in chronic sinusitis and allergen-induced nasal responses in seasonal allergic rhinitis. Am J Respir Cell Mol Biol 1997;17:683-90. 47. Ying S, Taborda-Barata L, Meng Q, Humbert M, Kay AB. The kinetics of allergen-induced transcription of messenger RNA for monocyte chemotactic protein-3 and RANTES in the skin of human atopic subjects: relationship to eosinophil, T cell, and macrophage recruitment. J Exp Med 1995;181:2153-9. 48. Reinecker H-C, Loh EY, Ringler DJ, Mehta A, Rombeau JL, MacDermott RP . Monocyte-chemoattractant protein 1 gene expression in intestinal epithelial cells and inflammatory bowel disease mucosa. Gastroenterology 1995;108:40-50. 49. Grimm MC, Doe WF. Chemokines in inflammatory bowel disease mucosa: expression of RANTES, macrophage inflammatory protein (MIP)-1a, MIP-1b, and ginterferon-inducible protein-10 by macrophages, lymphocytes, endothelial cells, and granulomas. Inflamm Bowel Dis 1996;2:88-96. 50. Gillitzer R, Wolff K, Tong D, et al. MCP-1 mRNA expression in basal keratinocytes of psoriatic lesions. J Invest Dermatol 1993;101:127-31. 51. Feng Y, Broder CC, Kennedy PE, Berger EA. HIV-1 entry cofactor: functional cDNA cloning of a seven-transmembrane, G protein-coupled receptor. Science 1996;272:872-7. 52. Dragic T, Litwin V, Allaway GP, et al. HIV-1 entry into CD4 cells is mediated by the chemokine receptor CC-CKR-5. Nature 1996;381:66773. 53. Deng H, Liu R, Ellmeier W, et al. Identification of a major co-receptor for primary isolates of HIV-1. Nature 1996;381:661-6. 54. Choe H, Farzan M, Sun Y, et al. The b-chemokine receptors CCR3 and CCR5 facilitate infection by primary HIV-1 isolates. Cell 1996;85: 1135-48. 55. Alkhatib G, Combadiere C, Broder CC, et al. CC CKR5: a RANTES, MIP-1a, MIP-1b receptor as a fusion cofactor for macrophage-tropic HIV1. Science 1996;272:1955-8. 56. Doranz BJ, Rucker J, Yi Y, et al. A dual-tropic primary HIV-1 isolate that uses fusin and the b-chemokine receptors CKR-5, CKR-3, and CKR2b as fusion cofactors. Cell 1996;85:1149-58. 57. DSouza MP , Harden VA. Chemokines and HIV-1 second receptors: confluence of two fields generates optimism in AIDS research. Nat Med 1996;2:1293-300. 58. Cocchi F, DeVico AL, Garzino-Demo A, Arya SK, Gallo RC, Lusso P . Identification of RANTES, MIP-1a, and MIP-1b as the major HIV-suppressive factors produced by CD8 T cells. Science 1995;270:1811-5. 59. Bleul CC, Farzan M, Choe H, et al. The lymphocyte chemoattractant SDF-1 is a ligand for LESTR/fusin and blocks HIV-1 entry. Nature 1996; 382:829-33. 60. Oberlin E, Amara A, Bachelerie F, et al. The CXC chemokine SDF-1 is the ligand for LESTR/fusin and prevents infection by T-cell-line-adapted HIV-1. Nature 1996;382:833-5. 61. Dean M, Carrington M, Winkler C, et al. Genetic restriction of HIV1 infection and progression to AIDS by a deletion allele of the CKR5 structural gene: Hemophilia Growth and Development Study, Multicenter AIDS Cohort Study, Multicenter Hemophilia Cohort Study, San Francisco City Cohort, ALIVE Study. Science 1996;273:1856-62. 62. Samson M, Libert F, Doranz BJ, et al. Resistance to HIV-1 infection in caucasian individuals bearing mutant alleles of the CCR-5 chemokine receptor gene. Nature 1996;332:722-5. 63. Liu R, Paxton WA, Choe S, et al. Homozygous defect in HIV-1 coreceptor accounts for resistance of some multiply-exposed individuals to HIV-1 infection. Cell 1996;86:367-77. 64. Paxton WA, Martin SR, Tse D, et al. Relative resistance to HIV-1 of

CD4 lymphocytes from persons who remain uninfected despite multiple high-risk sexual exposure. Nat Med 1996;2:412-7. 65. Wu L, Gerard NP, Wyatt R, et al. CD4-induced interaction of primary HIV-1 gp120 glycoproteins with the chemokine receptor CCR-5. Nature 1996;384:179-83. 66. Trkola A, Dragic T, Arthos J, et al. CD4-dependent, antibody-sensitive interactions between HIV-1 and its co-receptor CCR-5. Nature 1996;384: 184-7. 67. Lapham CK, Ouyang J, Chandrasekhar B, Nguyen NY, Dimitrov DS, Golding H. Evidence of cell-surface association between fusin and the CD4-gp120 complex in human cell lines. Science 1996;274:602-5. 68. Arvanitakis L, Geras-Raaka E, Varma A, Gershengorn MC, Cesarman E. Human herpesvirus KSHV encodes a constitutively active G-proteincoupled receptor linked to cell proliferation. Nature 1997;385:347-50. 69. Sharpe RJ, Byers HR, Scott CF, Bauer SI, Maione TE. Growth inhibition of murine melanoma and human colon carcinoma by recombinant human platelet factor 4. J Natl Cancer Inst 1990;82:848-53. 70. Luster AD, Leder P . IP-10, a -C-X-C- chemokine, elicits a potent thymus-dependent antitumor response in vivo. J Exp Med 1993;178:105765. 71. Strieter RM, Polverini PJ, Arenberg DA, et al. Role of C-X-C chemokines as regulators of angiogenesis in lung cancer. J Leukoc Biol 1995;57: 752-62. 72. Koch AE, Polverini PJ, Kunkel SL, et al. Interleukin-8 as a macrophage-derived mediator of angiogenesis. Science 1992;258:1798-801. 73. Graham GJ, Wright EG, Hewick R, et al. Identification and characterization of an inhibitor of haemopoietic stem cell proliferation. Nature 1990;344:442-4. 74. Van Damme J, Proost P, Lenaerts J-P, Opdenakker G. Structural and functional identification of two human, tumor-derived monocyte chemotactic proteins (MCP-2 and MCP-3) belonging to the chemokine family. J Exp Med 1992;176:59-65. 75. Streiter RM, Standiford TJ, Huffnagle GB, Colletti LM, Lukacs NW, Kunkel SL. The good, the bad, and the ugly: the role of chemokines in models of human disease. J Immunol 1996;156:3583-6. 76. Sekido N, Mukaida N, Harada A, Nakanishi I, Watanabe Y, Matsushima K. Prevention of lung reperfusion injury in rabbits by a monoclonal antibody against interleukin-8. Nature 1993;365:654-7. 77. Cook DN, Beck MA, Coffman TM, et al. Requirement of MIP-1a for an inflammatory response to viral infection. Science 1995;269:1583-5. 78. MacLean JA, Ownbey R, Luster AD. T cell-dependent regulation of eotaxin in antigen-induced pulmonary eosinophilia. J Exp Med 1996;184: 1461-9. 79. Sarafi MN, Garcia-Zepeda EA, MacLean JA, Charo IF, Luster AD. Murine monocyte chemoattractant protein (MCP)-5: a novel CC chemokine that is a structural and functional homologue of human MCP-1. J Exp Med 1997;185:99-109. 80. Stafford S, Li H, Forsythe PA, Ryan M, Bravo R, Alam R. Monocyte chemotactic protein-3 (MCP-3)/fibroblast-induced cytokine (FIC) in eosinophilic inflammation of the airways and the inhibitory effects of an antiMCP-3/FIC antibody. J Immunol 1997;158:4953-60. 81. Gonzalo J-A, Lloyd CM, Kremer L, et al. Eosinophil recruitment to the lung in a murine model of allergic inflammation: the role of T cells, chemokines, and adhesion receptors. J Clin Invest 1996;98:2332-45. 82. Lukacs NW, Strieter RM, Warmington K, Lincoln P, Chensue SW, Kunkel SL. Differential recruitment of leukocyte populations and alteration of airway hyperreactivity by C-C family chemokines in allergic airway inflammation. J Immunol 1997;158:4398-404. 83. Wada T, Tomosugi N, Naito T, et al. Prevention of proteinuria by the administration of anti-interleukin 8 antibody in experimental acute immune complex-induced glomerulonephritis. J Exp Med 1994;180:113540. 84. Dilloo D, Bacon K, Holden W, et al. Combined chemokine and cytokine gene transfer enhances antitumor immunity. Nat Med 1996;2:10905.

Vo l u m e 3 3 8 Downloaded from www.nejm.org at MCDANIEL COLLEGE on May 07, 2004. Copyright 1998 Massachusetts Medical Society. All rights reserved.

Nu m b e r 7

445

You might also like