Download Cytokine function of heat shock proteins - AJP

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Macrophage wikipedia , lookup

Immune system wikipedia , lookup

Adaptive immune system wikipedia , lookup

Monoclonal antibody wikipedia , lookup

DNA vaccination wikipedia , lookup

Adoptive cell transfer wikipedia , lookup

Polyclonal B cell response wikipedia , lookup

Immunomics wikipedia , lookup

Molecular mimicry wikipedia , lookup

Immunosuppressive drug wikipedia , lookup

Cancer immunotherapy wikipedia , lookup

Innate immune system wikipedia , lookup

Psychoneuroimmunology wikipedia , lookup

Transcript
Am J Physiol Cell Physiol 286: C739–C744, 2004;
10.1152/ajpcell.00364.2003.
Invited Review
Cytokine function of heat shock proteins
Min-Fu Tsan1,2 and Baochong Gao1
1
Research Service, Veterans Affairs Medical Center, Washington 20422; and 2Department
of Medicine, Georgetown University, Washington, District of Columbia 20007
tumor necrosis factor-␣; lipopolysaccharide; macrophages; innate immune system
HEAT SHOCK PROTEINS (HSPs) are the most phylogenetically
conserved proteins present in all prokaryotes and eukaryotes
(20, 27, 45). Traditionally, HSPs are regarded as intracellular
molecules. With the availability of recombinant bacterial and
human HSPs, there has been an intense interest in the extracellular function of HSPs in recent years. It has been shown
that HSPs are potent activators of the innate immune system,
capable of inducing proinflammatory cytokine production by
the monocyte-macrophage system, and the activation and maturation of dendritic cells (antigen-presenting cells) (5, 17, 21,
38, 52, 81–83, 85). The purpose of this focused review is to
critically evaluate the reported cytokine function of HSPs, with
particular emphasis on the question of whether the reported
cytokine effects are in fact due to HSPs or to the contaminant(s) that are present in the HSP preparations.
HEAT SHOCK PROTEINS
HSPs are expressed both constitutively (cognate proteins)
and under stressful conditions (inducible forms). In addition to
heat shock, a variety of stressful situations including environmental (ultraviolet radiation or heavy metals), pathological
(infections or malignancies), or physiological (growth factors
or cell differentiation) stimuli induce a marked increase in HSP
synthesis, known as the stress response (33, 45). Upon necrotic
cell death, HSPs are leaked into the extracellular compartment
(10). In addition, HSPs can be released extracellularly independent of necrotic cell death in response to a number of
stressful conditions (8, 29, 44). The mechanism and the physAddress for reprint requests and other correspondence: M.-F. Tsan, MidAtlantic Regional Office (10R), Office of Research Oversight, Dept. of
Veterans Affairs, 50 Irving St. NW, Washington, DC 20422 (E-mail: [email protected]).
http://www.ajpcell.org
iological significance of the HSP release are not clear. However, HSPs are present in circulation of normal individuals (57,
87), and their circulating levels are decreased in aging (62) and
increased in a number of pathological conditions such as
hypertension (58), atherosclerosis (87, 89), and after openheart surgery (18).
The primary function of the HSPs appears to serve as
molecular chaperones in which they recognize and bind to
nascent polypeptide chains and partially folded intermediates
of proteins, preventing their aggregation and misfolding, or as
chaperonins that directly mediate protein folding (20, 27, 33,
45). The classification of HSPs is based on their related
function and size (molecular mass). Using the nomenclature
adopted after the Cold Spring Harbor Meeting of 1996 (30),
family names are written in uppercase, e.g., HSP70, whereas
members of a family are conventionally written as Hsps, e.g.,
Hsp70.
Major classes of HSPs include the small HSPs, HSP40, 60,
70, 90, and 110 families (20, 27, 45). In mammalian species,
the HSP60 (chaperonin) family consists of mitochondrial
Hsp60 (mt-Hsp60) and cytosolic Hsp60 (T-complex polypeptide-1) (20, 27, 45). The mt-Hsp60 exists in a dynamic equilibrium among monomers, heptamers, and tetradecamers (20,
41). It dissociates into monomers at low concentrations and
assembles into tetradecamers in the presence of ATP and
mt-Hsp10, the cofactor of mt-Hsp60 (42). The cytosolic Hsp60
forms heterooligomeric ring structures and functions in cytosol
to fold cytoskeletal proteins such as actin and tubulin (46). The
HSP70 family includes the constitutive cytosolic Hsc70 (or
Hsp73), the stress-induced cytosolic Hsp70 (or Hsp72), the
endoplasmic reticulum (ER) Bip (or Grp78), and the mitochondrial mt-Hsp70 (20, 27, 45). The Hsp70 is composed of two
major functional domains. The NH2-terminal, highly conC739
Downloaded from http://ajpcell.physiology.org/ by 10.220.33.6 on April 28, 2017
Tsan, Min-Fu, and Baochong Gao. Cytokine function of heat shock proteins. Am
J Physiol Cell Physiol 286: C739–C744, 2004; 10.1152/ajpcell.00364.2003.—
Extensive work in the last 10 years has suggested that heat shock proteins (HSPs)
may be potent activators of the innate immune system. It has been reported that
Hsp60, Hsp70, Hsp90, and gp96 are capable of inducing the production of
proinflammatory cytokines by the monocyte-macrophage system and the activation
and maturation of dendritic cells (antigen-presenting cells) in a manner similar to
the effects of lipopolysaccharide (LPS) and bacterial lipoprotein, e.g., via CD14/
Toll-like receptor2 (TLR2) and CD14/TLR4 receptor complex-mediated signal
transduction pathways. However, recent evidence suggests that the reported cytokine effects of HSPs may be due to the contaminating LPS and LPS-associated
molecules. The reasons for previous failure to recognize the contaminant(s) as
being responsible for the reported HSP cytokine effects include failure to use highly
purified, low-LPS preparations of HSPs; failure to recognize the heat sensitivity of
LPS; and failure to consider contaminant(s) other than LPS. Thus it is essential that
efforts should be directed to conclusively determine whether the reported HSP
cytokine effects are due to HSPs or to contaminant(s) present in the HSP preparations before further exploring the implication and therapeutic potential of the
putative cytokine function of HSPs.
Invited Review
C740
HSP AND CYTOKINE FUNCTION
HSPS AND CYTOKINE FUNCTION
The above-described molecular chaperone function and presentation of antigens depend on the peptide-binding properties
of HSPs. Recent studies suggest that HSPs may also have
potent cytokine-like function independent of peptide binding.
HSPs such as Hsp60, Hsp70, Hsp90, and gp96 from a
variety of sources, including purified preparations from bacterial (23, 65, 69) and mammalian (4, 10, 52, 68, 71, 83) sources
as well as recombinant bacterial (15, 22, 37, 39, 53, 54, 66, 81,
86, 91) and human (5, 6, 13, 17, 18, 37, 38, 51, 81, 82)
products, have been shown to be potent activators of the innate
immune system (85). These HSP preparations have been
shown to induce the production of proinflammatory cytokines
such as tumor necosis factor (TNF)-␣, interleukin (IL)-1, IL-6,
and IL-12 and the release of nitric oxide (NO) and C-C
chemokines by monocytes, macrophages, and dendritic cells.
They also induce the maturation of dendritic cells as demonstrated by the upregulation of MHC class I and II molecules,
CD86, CD40, etc. (10, 21, 52, 68, 71, 86). The Hsp60 and
Hsp70 preparations purified from bacterial sources or from
recombinant bacterial and human products are capable of
inducing the above effects in concentrations ranging from ⬍1
AJP-Cell Physiol • VOL
␮g/ml to a few micrograms per milliliter, whereas Hsp70,
Hsp90, and gp96 isolated from mouse liver require concentrations that are one to two orders of magnitude higher (e.g.,
10–100 ␮g/ml). These HSP cytokine effects, compared with
their molecular chaperone function, are unique in that they
require no HSP-associated peptides, no ATP hydrolysis, no
cofactors, and no protein complex assembly. A new term,
“chaperokine,” has been coined for HSPs to indicate their dual
functions as molecular chaperones and cytokines (5).
Furthermore, the observed HSP cytokine effects are mediated via the CD14/Toll-like receptor (both TLR2 and TLR4)
complex signal transduction pathways leading to the activation
of nuclear factor-␬B (NF-␬B) and mitogen-activated protein
kinases (MAPKs), i.e., ERKs (p42 and p44 extracellular signal-regulated kinases), JNK (c-Jun NH2-terminal kinase), and
p38 kinase (5, 6, 15, 18, 38, 51, 81–83). The CD14 and TLR
receptor complexes are pattern recognition receptors involved
in the innate immunity for the pathogen recognition and host
defense (3, 47). CD14, the lipopolysaccharide (LPS) receptor,
is a glycophosphatidyl inositol (GPI)-anchored membrane protein lacking transmembrane and intracellular domains (28, 76).
TLRs are type I transmembrane proteins with an extracellular
domain containing a leucine-rich repeat and a cytoplasmic
domain analogous to that of the IL-1 receptor (IL-1R) family
(47, 67). An adapter protein, MyD88 (myeloid differentiation
protein 88), binds to the Toll/IL-1R homology (TIR) motif
through its own TIR motif, whereas a death domain on its
COOH terminus recruits IL-1R-associated kinase (IRAK) to
the complex (48). IRAK is then autophosphorylated and released from the complex to bind TRAF6 (TNF receptorassociated factor 6), which can then activate either the NF-␬B
or the MAPKs (47, 49). Together with CD14 and an accessory
protein MD2, TLR4 initiates signaling cascades in response to
LPS, whereas TLR2 initiates the signal cascades in response to
bacterial lipoprotein, Gram-positive bacteria, yeast, and spirochetes (2, 14, 34, 47, 67).
The reported activation of the innate immune system by
HSPs, as described above, has been hailed as an important new
function of HSPs with broad biological significance. The
induction of proinflammatory cytokines by Hsp60 and Hsp70
may contribute to the pathogenesis of autoimmune diseases
and chronic inflammation (55). Chlamydial Hsp60 frequently
colocalizes with human Hsp60 in macrophages of atherosclerotic plaques (39). Induction of proinflammatory cytokine
release from macrophages by chlamydial Hsp60 would provide
a potential mechanism by which chlamydial infections may
promote atherogenesis and precipitate acute ischemic events
(37, 39). Likewise, the activation and maturation of dendritic
cells by gp96 may be responsible for the gp96-induced tumor
immunity by inducing both the innate and adaptive immune
responses (50). Thus it has been proposed that through their
cytokine function, HSPs may serve as a “danger signal” to
the innate immune system at the site of tissue injury (17, 85)
and that HSPs could be the endogenous ligands for the
TLR2 and TLR4 (51, 82). In fact, HSPs are considered to be
the prototype of endogenous ligands for Toll-like receptors
(12). There is considerable interest to further explore the
implications and therapeutic potential of these HSP cytokine
effects (7, 55, 88).
286 • APRIL 2004 •
www.ajpcell.org
Downloaded from http://ajpcell.physiology.org/ by 10.220.33.6 on April 28, 2017
served ATPase domain binds ADP and ATP tightly and hydrolyzes ATP, whereas the COOH-terminal domain is required
for polypeptide binding (20, 27). The HSP90 family includes
the cytosolic Hsp90 (␣ and ␤) and the ER form, gp96 (grp94).
Glucose-regulated proteins (grp) such as grp78 and grp94/gp96
are molecular chaperones in the ER that are upregulated in
response to glucose starvation and other stressful stimuli that
disrupt protein folding in the ER (30, 70). For more detailed
description of HSPs, we refer the reader to excellent reviews of
the subject (20, 27, 33, 45, 55, 70, 87).
Bacterial HSPs, particularly Hsp60 and Hsp70, are highly
immunogenic, capable of inducing antibody production and
T-cell activation (92). The antibodies and T cells against
bacterial Hsp60 and Hsp70 also recognize mammalian Hsp60
and Hsp70, respectively, due to cross-reactivity (36). These
anti-Hsp60 and anti-Hsp70 antibodies and T cells injure tissues
and cause inflammatory reactions. Thus Hsp60 and Hsp70
have been implicated in the pathogenesis of a number of
autoimmune diseases and inflammatory conditions such as type
1 diabetes (1, 19), Crohn’s disease (77), atherosclerosis (56,
87), and juvenile chronic arthritis (60, 64).
HSPs also play important roles in antigen presentation,
cross-presentation, and tumor immunity (43, 72, 73). HSPs of
the cytosol, such as Hsp70 and Hsp90, and of the ER, such as
gp96, bind antigenic peptides generated within the cells and are
part of the endogenous pathway of antigen presentation by the
major histocompatibility complex (MHC) class I molecules
(32, 43, 74). Peptides that are chaperoned by HSPs when
released extracellularly are taken up by antigen-presenting
cells via ␣2-macroglobulin receptor (CD91)-mediated endocytosis, resulting in representation by the MHC molecules (9, 43,
75). Vaccination of mice with Hsp70, Hsp90, and gp96 isolated from murine tumor cells elicits immune response sufficient for tumor rejection and suppression of metastatic tumor
progression (73). This tumor immunity results from tumorderived peptides associated with the HSPs, rather than from the
HSPs themselves (80).
Invited Review
C741
HSP AND CYTOKINE FUNCTION
HSPS VS. CONTAMINANTS
Fig. 1. Effect of heat inactivation on endotoxin
activity and tumor necrosis factor (TNF)-␣-inducing
activity of lipopolysaccharide (LPS). A stock solution of LPS at 4 ng/ml was heated in a boiling water
bath for 1 h. A: endotoxin activities of the nonheated LPS and heated LPS were determined using
the Limulus amebocyte lysate assay. B: murine macrophages were treated with LPS or heated LPS at the
indicated concentrations for 4 h. TNF-␣ concentrations in media were then determined. Values represent means ⫾ SD of 3 experiments. *P ⬍ 0.05 vs.
non-heated LPS. [Reprinted from Gao and Tsan (25)
with permission.]
AJP-Cell Physiol • VOL
286 • APRIL 2004 •
www.ajpcell.org
Downloaded from http://ajpcell.physiology.org/ by 10.220.33.6 on April 28, 2017
The reported HSP cytokine effects are similar to those of
LPS and bacterial lipoprotein. Because the recombinant bacterial and human HSPs are produced by Escherichia coli
expressing HSP cDNAs, the final preparations may be contaminated with bacterial products. Likewise, HSP preparations
isolated from bacteria or murine tissues are also frequently
contaminated with LPS. The fact that various HSPs all have
similar effects and share the same CD14/TLR2 and CD14/
TLR4 receptor complexes is of considerable concern.
Ample examples exist in the literature demonstrating how
contaminants can lead to misleading conclusions. For example,
in 1998 with the use of the commercially available LPS
preparation, it was first reported (90) that TLR2 mediated the
LPS-induced activation of NF-␬B and could be the long sought
after LPS signal transducer. There followed a period of uncertainty regarding whether TLR2 or TLR4 was the LPS signal
transducer (35, 59, 61). In 2000, it was then demonstrated that
TLR2 could not mediate cellular response by repurified commercial preparations of LPS (31, 78). In 2002, two lipoproteins
(Lip12 and Lip19) extracted from E. coli LCD25 LPS were
identified to be the major components responsible for TLR2mediated cell activation in the commercial LPS preparations
(40). Thus failure to recognize the presence of lipoproteins in
the commercially available LPS preparation led to the erroneous attribution of lipoprotein signal transducer, TLR2, as the
LPS signal transducer (35, 90).
Investigators are cognizant of the possibility of contamination, particularly LPS, and have attempted to rule out the
possibility of LPS contamination being responsible for the
observed HSP cytokine effects. Most studies have used two
criteria: first, LPS is resistant to heat inactivation (5, 17, 18, 23,
38, 65, 66, 86), and second, LPS effects are inhibitable by
polymyxin B (5, 17, 18, 38, 65, 66, 86). Other less frequently
used criteria include the effect of anti-HSP antibodies (15, 65)
and other LPS inhibitors such as lipid IVa (5), LPS (10, 52), or
lipid A (18) from Rhodopseudomonas spheroides. Because the
observed HSP cytokine effects were heat sensitive, either not
inhibitable or only partially inhibitable by polymyxin B, not
inhibitable by other LPS inhibitors, and/or inhibitable by antiHSP antibodies, it was concluded that the observed HSP
cytokine effects could not have been due to LPS contamination. However, doubts about these criteria have been raised.
Wallin et al. (85) noted that highly purified murine liver Hsp70
had no cytokine effects even at concentrations as high as
200–300 ␮g/ml. On the other hand, a LPS-contaminated prep-
aration at Hsp70 concentrations as low as 50–100 ng/ml caused
cytokine effects that were heat sensitive and were not inhibitable by polymyxin B.
Recent studies in which HSP preparations essentially free of
LPS were used suggest that the previously reported cytokine
function of HSPs may be due to the contaminants. Bausinger et
al. (11) reported that LPS-free recombinant human Hsp70
(rhHsp70) did not induce the activation of dendritic cells. Gao
and Tsan (24, 25) demonstrated that LPS was heat sensitive
(Fig. 1) and that the ability of commercially available rhHsp70
to induce TNF-␣ production was entirely due to the contaminating LPS (24), whereas that of rhHSP60 was due to contamination by LPS as well as LPS-associated molecules (25). Reed
et al. (63) reported that the activation of NF-␬B and the
production of NO by gp96 were due to LPS contamination.
Importantly, all of these investigators demonstrated that these
highly purified, essentially LPS-free HSPs retained their normal molecular chaperone functions or ATPase activity (11, 24,
25, 63). Thus failure of Hsp60, Hsp70, and gp96 to induce
cytokine or NO production by macrophages or to activate
antigen-presenting cells was not due to defective HSPs as a
result of purification.
The fact that LPS is sensitive to heat inactivation (24, 25, 84)
has not been widely appreciated. Most investigators are not
aware that macrophages are extremely sensitive to LPS. LPS at
a concentration of 0.1–0.2 ng/ml is sufficient to maximally
induce TNF-␣ release from murine macrophages (24). However, in most studies LPS was used at concentrations ranging
from 10 to 500 ng/ml to test for heat sensitivity (5, 17, 18, 23,
38, 65, 66, 86). At these concentrations, even if heat treatment
inactivated 99% of the LPS used in the studies, there would
still be sufficient residual LPS to induce TNF-␣ release, giving
the impression that LPS was heat resistant. Thus, unless one
uses an LPS concentration similar to the LPS concentration
present in the HSP preparation, the result could be misleading.
Although LPS may be the most frequent contaminant, nonLPS contaminant(s) capable of inducing proinflammatory cytokines may also contribute to the reported cytokine effects of
HSPs. Gao and Tsan (25) showed that 50% of the TNF-␣inducing activity of the commercially available rhHsp60 (no.
NSP-540; StressGen Biotechnologies, Victoria, BC, Canada)
was due to non-LPS contaminant(s) that was heat sensitive but
not inhibitable by polymyxin B. The presence of non-LPS
contaminant(s) could partially explain previous reports that the
observed cytokine effects of HSPs were either not inhibitable
Invited Review
C742
HSP AND CYTOKINE FUNCTION
AREAS OF UNCERTAINTY
Although recent evidence suggests that the cytokine effects
that have been reported in the last 10 years may be due to the
effects of contaminants such as LPS and LPS-associated molecules (11, 24, 25, 63), a number of uncertainties exist.
It is not clear why anti-Hsp60 antibodies could inhibit
Hsp60-induced proinflammatory cytokine production by macrophages (15, 65). It is possible that HSPs bind LPS and that
binding of anti-HSP antibodies to HSPs interferes with the
interaction of HSP-bound LPS with the CD14/TLR4 receptor
complex, thus inhibiting the effect of LPS. There is supporting
evidence that Hsp70, Hsp90, and gp96 bind LPS (16, 63, 79).
If the reported cytokine effects are conclusively shown to be
due to contaminants, not due to HSPs themselves, then one
important question will be whether HSPs have any effect on
the innate immune system. Habich et al. (26) have described a
macrophage receptor for Hsp60 that is specific and saturable,
suggesting that Hsp60 may have some effect on macrophages
that is different from the reported cytokine effects mediated
through the CD14/TLR receptor complexes. Reed et al. (63)
reported that a highly purified, low-LPS preparation of gp96
was able to elicit a marked increase in ERK phosphorylation
but not the activation of p38 and JNK, an effect distinct from
that of LPS. Thus HSPs may activate macrophage signal
transduction independent of LPS.
In conclusion, extensive investigation in the past 10 years
has suggested that HSPs may be potent activators of the innate
immune system. It has been shown that Hsp60, Hsp70, Hsp90,
and gp96 are capable of inducing the production of proinflammatory cytokines by the monocyte-macrophage system, as well
as the activation and maturation of antigen-presenting cells in
a manner similar to the effects of LPS and bacterial lipoprotein,
e.g., via CD14/TLR2 and CD14/TLR4 receptor complex-mediated signal transduction pathways. However, recent evidence
suggests that the reported cytokine effects of HSPs may be due
to LPS and LPS-associated molecules. Thus it is essential that
efforts should be directed to conclusively determine whether
the reported HSP cytokine effects are due to HSPs or to
contaminant(s) present in the HSP preparations before exploring further the implication and therapeutic potential of the
putative cytokine function of HSPs.
AJP-Cell Physiol • VOL
GRANTS
This work is supported by the Medical Research Service, Office of Research and Development, Department of Veterans Affairs.
REFERENCES
1. Abulafia-Lapid R, Elias D, Raz L, Keren-Zur Y, Atlan H, and Cohen
IR. T cell proliferative responses of type 1 diabetes patients and healthy
individuals to human hsp60 and its peptides. J Autoimmun 12: 121–129,
1999.
2. Aliprantis AO, Yang RB, Mark MR, Suggett S, Devaux B, Radolf JD,
Klimpel GR, Godowski PJ, and Zychlinsky A. Cell activation and
apoptosis by bacterial lipoproteins through Toll-like receptor 2. Science
285: 736–739, 1999.
3. Antal-Szalmas P. Evaluation of CD14 in host defense. Eur J Clin Invest
30: 167–179, 2000.
4. Asea A, Kabingu E, Stevenson MA, and Calderwood SK. HSP70
peptide-bearing and peptide-negative preparations act as chaperokines.
Cell Stress Chaperones 5: 425–431, 2000.
5. Asea A, Kraeft SK, Kurt-Jones EA, Stevenson MA, Chen LB, Finberg
RW, Koo GC, and Calderwood SK. Hsp70 stimulates cytokine production through a CD-14-dependent pathway, demonstrating its dual role as a
chaperone and cytokine. Nat Med 6: 435–442, 2000.
6. Asea A, Rehli M, Kabingu E, Boch JA, Bare O, Auron PE, Stevenson
MA, and Calderwood SK. Novel signal transduction pathway utilized by
extracellular hsp70: role of toll-like receptor (TLR) 2 and TLR4. J Biol
Chem 277: 15028–15034, 2002.
7. Audibut F. Adjuvants for vaccine. A quest. Int Immunopharmacol 3:
1187–1193, 2003.
8. Barreto A, Gonzalez JM, Kabingu E, Asea A, and Fiorentino S.
Stress-induced release of HSC70 from human tumors. Cell Immunol 222:
97–104, 2003.
9. Basu S, Binder RJ, Ramalingam T, and Srivastava PK. CD91 is a
common receptor for heat shock proteins gp96, hsp90, hsp70, and calreticulin. Immunity 14: 303–313, 2001.
10. Basu S, Binder RJ, Suto R, Anderson KM, and Srivastava PK.
Necrotic but not apoptotic cell death releases heat shock proteins, which
deliver a partial maturation signal to dendritic cells and activate the NF-␬B
pathway. Int Immunol 12: 1539–1546, 2000.
11. Bausinger H, Lipsker D, Ziylan U, Manie S, Briand JP, Cazenave JP,
Muller S, Haeuw JF, Ravanat C, de la Salle H, and Hanau D.
Endotoxin-free heat shock protein 70 fail to induce APC activation. Eur
J Immunol 32: 3708–3713, 2002.
12. Beg AA. Endogenous ligands of Toll-like receptors: implications for
regulating inflammatory and immune responses. Trends Immunol 23:
509–512, 2002.
13. Bethke K, Staib F, Distler M, Schmitt U, Jonuleit H, End AH, Galle
PR, and Heike M. Different efficiency of heat shock proteins (HSP) to
activate human monocytes and dendritic cells: superiority of HSP60.
J Immunol 169: 6141–6148, 2002.
14. Beutler B. Tlr4: central component of the sole mammalian LPS sensor.
Curr Opin Immunol 12: 20–26, 2000.
15. Bulut Y, Faure E, Thomas L, Karahashi H, Michelsen KS, Equils O,
Morrison SG, Morrison RP, and Arditi M. Chlamydial heat shock
protein 60 activates macrophages and endothelial cells through Toll-like
receptor 4 and MD2 in a MyD88-dependent pathway. J Immunol 168:
1435–1440, 2002.
16. Byrd CA, Bornmann W, Erdjument-Bromage H, Trmpst P, Pavletich
N, Rosen N, Nathan CF, and Ding A. Heat shock protein 90 mediates
macrophage activation by taxol and bacterial lipopolysaccharide. Proc
Natl Acad Sci USA 96: 5645–5650, 1999.
17. Chen W, Syldath U, Bellmann K, Burkart V, and Kolb H. Human
60-kDa heat-shock protein: a danger signal to the innate immune system.
J Immunol 162: 3212–3219, 1999.
18. Dybdahl B, Wahba A, Lien E, Flo TH, Waage A, Qureshi N, Sellevold
OFM, Espevik T, and Sundan A. Inflammatory response after open heart
surgery: release of heat-shock protein 70 and signaling through Toll-like
receptor-4. Circulation 105: 685–690, 2002.
19. Elias D, Reshef T, Birk OS, van der Zee R, Walker MD, and Cohen
IR. Vaccination against autoimmune mouse diabetes with a T-cell epitope
of the human 65-kDa heat shock protein. Proc Natl Acad Sci USA 88:
3088–3091, 1991.
20. Fink AL. Chaperone-mediated protein folding. Physiol Rev 79: 425–449,
1999.
286 • APRIL 2004 •
www.ajpcell.org
Downloaded from http://ajpcell.physiology.org/ by 10.220.33.6 on April 28, 2017
or only partially inhibitable by polymyxin B (5, 17, 18, 38, 65,
66, 85, 86).
It is thus important to determine conclusively whether HSPs
can activate the innate immune system before further investigating the implication and therapeutic potential of the reported
cytokine effects of HSPs. Only highly purified HSP preparations that are essentially free of LPS contamination should be
used. Such low LPS preparations for rhHsp 60 and rhHsp70 are
commercially available. However, it is important to measure
the endotoxin activity of each preparation by using the Limulus
amebocyte lysate assay beforehand to ensure that the preparation is essentially free of LPS contamination. If the Hsp60 or
Hsp70 preparation is contaminated with LPS, then LPS and
LPS-associated proteins can be efficiently removed with the
use of commercially available polymyxin B agarose gel without affecting the chaperone functions of HSPs (11, 24, 25).
Because gp96 binds LPS tightly, the novel purification procedures as reported by Reed et al. (63) should be used.
Invited Review
HSP AND CYTOKINE FUNCTION
AJP-Cell Physiol • VOL
44. Liao DF, Jin ZG, Baas AS, Daum G, Gygi SP, Aebersold R, and Berk
BC. Purification and identification of secreted oxidative stress-induced
factors from vascular smooth muscle cells. J Biol Chem 275: 189–196,
2000.
45. Lindquist S and Craig EA. The heat-shock proteins. Annu Rev Genet 22:
631–677, 1988.
46. Llorca O, Martı́n-Benito J, Ritco-Vonsovici M, Grantham J, Hynes
GM, Willison KR, Carrascosa JL, and Valpuesta JM. Eukaryotic
chaperonin CCT stabilizes actin and tubulin folding intermediates in open
quasi-native conformations. EMBO J 19: 5971–5979, 2000.
47. Medzhitov R. Toll-like receptors and innate immunity. Nat Rev Immunol
1: 135–145, 2001.
48. Medzhitov R, Preston-Hurlburt P, Kopp E, Stadlen A, Chen C, Ghosh
S, and Janeway CA Jr. MyD88 is an adapter protein in the hToll/IL-1
receptor family signaling pathways. Mol Cell 2: 253–258, 1998.
49. Muzio M, Natoli G, Saccani S, Levrero M, and Mantovani A. The
human toll signaling pathway: divergence of nuclear factor kappaB and
JNK/SAPK activation upstream of tumor necrosis factor receptor-associated factor 6 (TRAF6). J Exp Med 187: 2097–2101, 1998.
50. Nicchitta CV. Re-evaluating the role of heat-shock protein-peptide interactions in tumor immunity. Nat Rev Immunol 3: 427–432, 2003.
51. Ohashi K, Burkart V, Flohe S, and Kolb H. Cutting edge: heat shock
protein 60 is a putative endogenous ligand of the toll-like receptor-4
complex. J Immunol 164: 558–561, 2000.
52. Panjwani NN, Popova L, and Srivastava PK. Heat shock protein gp96,
and hsp70 activate the release of nitric oxide by APCs. J Immunol 168:
2997–3003, 2002.
53. Peetermans WE, Raats CJI, Langemans JAM, and Van Furth R.
Mycobacterial heat-shock protein 65 induces proinflammatory cytokines
but does not activate human mononuclear phagocytes. Scand J Immunol
39: 613–617, 1994.
54. Peetermans WE, Raats CJI, van Furth R, and Langermans JAM.
Mycobacterial 65-kilodalton heat shock protein induces tumor necrosis
factor ␣, interleukin 6, reactive nitrogen intermediates and toxoplasmastatic activity in murine peritoneal macrophages. Infect Immun 63: 3454–
3458, 1995.
55. Pockley AG. Heat shock proteins in health and disease: therapeutic targets
or therapeutic agents? In: Expert Reviews in Molecular Medicine. Cambridge, UK: Cambridge University Press, 2001, p. 1–21.
56. Pockley AG. Heat shock proteins, inflammation, and cardiovascular
disease. Circulation 105: 1012–1017, 2002.
57. Pockley AG, Bulmer J, Hanks BM, and Wright BH. Identification of
human heat shock protein 60 (Hsp60) and anti-Hsp60 antibodies in the
peripheral circulation of normal individuals. Cell Stress Chaperones 4:
29–35, 1999.
58. Pockley AG, Wu R, Lemne C, Kiessling R, de Faire U, and Frostegard
J. Circulating heat shock protein 60 is associated with early cardiovascular
disease. Hypertension 36: 303–307, 2000.
59. Poltorak A, He X, Smirnova I, Liu MY, van Huffel C, Du X, Birdwell
D, Alejos E, Silva M, Galanos C, Freudenberg M, Ricciardi-Castagnoli P, Layton B, and Beutler B. Defective LPS signaling in C3H/HeJ
and C57BL/10ScCr mice: mutations in Tlr4 gene. Science 282: 2085–
2088, 1998.
60. Pope RM, Lovis RM, and Gupta RS. Activation of synovial fluid T
lymphocytes by 60-kd heat shock proteins in patients with inflammatory
synovitis. Arthritis Rheum 35: 43–48, 1992.
61. Qureshi ST, Lariviere L, Levegue G, Clermont SKJ, Moore Gros P,
and Malo D. Endotoxin-tolerant mice have mutations in Toll-like receptor
4 (Tlr4). J Exp Med 189: 615–625, 1999.
62. Rea IM, McNerlan S, and Pockley AG. Serum heat shock protein and
anti-heat shock protein antibody levels in aging. Exp Gerontol 36: 341–
352, 2001.
63. Reed RC, Berwin B, Baker JP, and Nicchitta CV. GRP94/gp96 elicits
ERK activation in murine macrophages: a role for endotoxin contamination in NF␬B activation and nitric oxide production. J Biol Chem 278:
31853–31860, 2003.
64. Res PC, Schear CG, Breedveld FC, van Eden W, Van Embden JD,
Cohen IR, and De Vries RR. Synovial fluid T cell reactivity against 65
kD heat shock protein of mycobacteria in early chronic arthritis. Lancet 2:
478–480, 1988.
65. Retzlaff C, Yamamoto Y, Hoffman PS, Friedman H, and Klein TW.
Bacterial heat shock proteins directly induce cytokine mRNA and interleukin-1 secretion in macrophage cultures. Infect Immun 62: 5689–5693,
1994.
286 • APRIL 2004 •
www.ajpcell.org
Downloaded from http://ajpcell.physiology.org/ by 10.220.33.6 on April 28, 2017
21. Flohe SB, Bruggemann J, Lendemans S, Nikulina M, Meierhoff G,
Flohe S, and Kolb H. Human heat shock protein 60 induces maturation
of dendritic cells versus a TH1-promoting phenotype. J Immunol 170:
2340–2348, 2003.
22. Friedland JS, Shattock R, Remake DG, and Griffin E. Mycobacterial
65-kd heat shock protein induces release of proinflammatory cytokines
from human monocytic cells. Clin Exp Immunol 91: 58–62, 1993.
23. Galdiero M, DeL’ero GC, and Marcatili A. Cytokine and adhesion
molecule expression in human monocytes and endothelial cells stimulated
with bacterial heat shock proteins. Infect Immun 65: 699–707, 1997.
24. Gao B and Tsan MF. Endotoxin contamination in recombinant human
Hsp70 preparation is responsible for the induction of TNF␣ release by
murine macrophages. J Biol Chem 278: 174–179, 2003.
25. Gao B and Tsan MF. Recombinant human heat shock protein 60 does not
induce the release of tumor necrosis factor ␣ from murine macrophages.
J Biol Chem 278: 22523–22529, 2003.
26. Habich C, Baumgart K, Kolb H, and Burkart V. The receptor for heat
shock protein 60 on macrophages is saturable, specific, and distinct from
receptors for other heat shock proteins. J Immunol 168: 569–576, 2002.
27. Hartl FU and Hayer-Hartl M. Molecular chaperones in the cytosol: from
nascent chain to folded protein. Science 295: 1852–1858, 2002.
28. Haziot A, Chen S, Ferrero E, Low MG, Silber R, and Goyert SM. The
monocyte differentiation antigen, CD14, is anchored to the cell membrane
by a phosphatidylinositol linkage. J Immunol 141: 547–552, 1988.
29. Hightower LE and Guidon PT Jr. Selective release from cultured
mammalian cells of heat-shock (stress) proteins that resemble glia-axon
transfer proteins. J Cell Physiol 138: 257–266, 1989.
30. Hightower LE and Hendershot LM. Molecular chaperones and the heat
shock response at Cold Spring Harbor. Cell Stress Chaperones 2: 1–11,
1997.
31. Hirschfeld M, Ma Y, Weis JH, Vogel SN, and Weis JJ. Cutting edge:
repurification of lipopolysaccharide eliminates signaling through both
human and murine Toll-like receptor 2. J Immunol 165: 618–622, 2000.
32. Ishii T, Udono H, Ohta H, Ono T, Hizuta A, Tanaka N, Srivastava PK,
and Nakayama E. Isolation of MHC class 1-restricted antigen peptide
and its precursors associated with heat shock proteins hsp70, hsp90, and
gp96. J Immunol 162: 1303–1309, 1999.
33. Jaattela M. Heat shock proteins as cellular lifeguards. Ann Med 31:
261–271, 1999.
34. Jones BW, Heldwein KA, Means TK, Saukkonen JJ, and Fenton MJ.
Differential roles of Toll-like receptors in the elicitation of proinflammatory responses by macrophages. Ann Rheum Dis 60, Suppl 3: iii6–iii12,
2001.
35. Kirschning CJ, Wesche H, Ayres TM, and Rothe M. Human Toll-like
receptor 2 confers responsiveness to bacterial lipopolysaccharide. J Exp
Med 188: 2091–2097, 1999.
36. Kissling R, Growbug A, Tvanyi J, Soderstrom K, Ferm M, Kleinau S,
Nisseon E, and Klareskog L. Role of hsp60 during autoimmune and
bacterial inflammation. Immunol Rev 121: 91–111, 1991.
37. Kol A, Bourcier T, Lichtman AH, and Libby P. Chlamydial and human
heat shock protein 60s activate human vascular endothelium, smooth
muscle cells, and macrophages. J Clin Invest 103: 571–577, 1999.
38. Kol A, Lichtman AH, Finberg RW, Libby P, and Kurt-Jones EA.
Cutting edge: heat shock protein (HSP) 60 activates the innate immune
response: CD14 is an essential receptor for HSP60 activation of mononuclear cells. J Immunol 164: 13–17, 2000.
39. Kol A, Sukhova GK, Lichtman AH, and Libby P. Chlamydial heat
shock protein 60 localizes in human atheroma and regulates macrophage
tumor necrosis factor-␣ and matrix metalloproteinase expression. Circulation 98: 300–307, 1998.
40. Lee HK, Lee J, and Tobias PS. Two lipoproteins extracted from
Escherichia coli K-12 LCD25 lipopolysaccharide are the major components responsible for Toll-like receptor 2-mediated signaling. J Immunol
168: 4012–4017, 2002.
41. Levy-Rimler G, Bell RE, Ben-Tal N, and Azem A. Type I chaperonins:
not all are created equal. FEBS Lett 529: 1–5, 2002.
42. Levy-Rimler G, Viitanen P, Weiss C, Sharkia R, Greenberg A, Niv A,
Lustig A, Delarea Y, and Azem A. The effect of nucleotides and
mitochondrial chaperonin 10 on the structure and chaperone activity of
mitochondrial chaperonin 60. Eur J Biochem 268: 3465–3472, 2001.
43. Li Z, Menoret A, and Srivastava P. Roles of heat-shock proteins in
antigen presentation and cross-presentation. Curr Opin Immunol 14:
45–51, 2002.
C743
Invited Review
C744
HSP AND CYTOKINE FUNCTION
AJP-Cell Physiol • VOL
80. Udono H and Srivastava PK. Comparison of tumor-specific immunogenicities of stress-induced proteins gp96, hsp90 and hsp70. J Immunol 152:
5398–5403, 1994.
81. Vabulas RM, Ahmad-Nejad P, da Costa C, Miethke T, Kirschning CJ,
Hacker H, and Wagner H. Endocytosed HSP60s use Toll-like receptor
2 (TLR2) and TLR4 to activate the Toll/interleukin-1 receptor signaling
pathway in innate immune cells. J Biol Chem 276: 31332–31339, 2001.
82. Vabulas RM, Ahmad-Nejad P, Ghose S, Kirschning CJ, Issels RD,
and Wagner H. Hsp70 as endogenous stimulus of the Toll/interleukin-1
receptor signal pathway. J Biol Chem 277: 15107–15112, 2002.
83. Vabulas RM, Braedel S, Hilf N, Singh-Jasuja H, Herter S, AhmadNejad P, Kirschning CJ, Da Costa C, Rammensee HG, Wagner H,
and Schild H. The endoplasmic reticulum-resident heat shock protein
Gp96 activates dendritic cells via the Toll-like receptor 2/4 pathway.
J Biol Chem 277: 20847–20853, 2002.
84. Vikstrom EV. The immunosuppressive activity of chemically modified
lipopolysaccharide of Shigella sonnei. Immunol Lett 80: 15–19, 2002.
85. Wallin RP, Lundqvist A, More SH, von Bonin A, Kiessling R, and
Ljunggren HG. Heat-shock proteins as activators of the innate immune
system. Trends Immunol 23: 130–135, 2002.
86. Wang Y, Kelly CG, Singh M, McGowan EG, Carrara AS, Bergmeier
LA, and Lehner T. Stimulation of Th-1 polarizing cytokines, C-C
chemokines, maturation of dendritic cells, and adjuvant function by the
peptide binding fragment of heat shock protein 70. J Immunol 169:
2422–2429, 2002.
87. Xu Q. Role of heat shock proteins in atherosclerosis. Arterioscler Thromb
Vasc Biol 22: 1547–1559, 2002.
88. Xu Q. Infections, heat shock proteins, and atherosclerosis. Curr Opin
Cardiol 18: 245–252, 2003.
89. Xu Q, Schett G, Perschinka H, Mayr M, Egger G, Oberhollenzer F,
Willeit J, Kiechl S, and Wick G. Serum soluble heat shock protein 60 is
elevated in subjects with atherosclerosis in a general population. Circulation 102: 14–20, 2000.
90. Yang RB, Mark MR, Gray A, Hunag A, Xie MH, Zhang M, Goddard
A, Wood WI, Gurney AL, and Godowski PJ. Toll-like receptor-2
mediates lipopolysaccharide-induced cellular signaling. Nature 395: 284–
288, 1998.
91. Zhang Y, Doerfler M, Lee TC, Guillemin B, and Rom WN. Mechanisms of stimulation of interleukin-1␤ and tumor necrosis factor ␣ by
Mycobacterium tuberculosis components. J Clin Invest 91: 2076–2083,
1993.
92. Zugel U and Kaufmann SH. Immune response against heat-shock
proteins in infectious diseases. Immunobiology 201: 22–35, 1999.
286 • APRIL 2004 •
www.ajpcell.org
Downloaded from http://ajpcell.physiology.org/ by 10.220.33.6 on April 28, 2017
66. Retzlaff C, Yamamoto Y, Okubo S, Hoffman PS, Friedman PS, and
Klein TW. Legionella pneumophila heat-shock protein-induced increase
of interleukin-1␤ mRNA involves protein kinase C signaling in macrophages. Immunology 89: 281–288, 1996.
67. Rock FL, Hardiman G, Timans JC, Kastelein RA, and Bazan JF. A
family of human receptors structurally related to Dorsophila Toll. Proc
Natl Acad Sci USA 95: 588–593, 1998.
68. Singh-Jasuja H, Scherer HU, Hilf N, Arnold-Schild D, Rammensee
HG, Toes REM, and Schild H. The heat shock protein gp96 induces
maturation of dendritic cells and down-regulation of its receptor. Eur
J Immunol 30: 2211–2215, 2000.
69. Skeen MJ, Miller MA, Shinnick TM, and Ziegler HK. Regulation of
murine macrophage IL-12 production. Activation of macrophages in vivo,
restimulation in vitro, and modulation by other cytokines. J Immunol 156:
1196–1206, 1996.
70. Snoeckx LHEH, Cornelussen RN, Van Nieuwenhoven FA, Reneman
RS, and Van Der Vusse GJ. Heat shock proteins and cardiovascular
pathophysiology. Physiol Rev 81: 1461–1497, 2001.
71. Somersan S, Larsson M, Honteneau JF, Basu S, Srivastava P, and
Bhardwaj N. Primary tumor tissue lysates are enriched in heat shock
proteins and induce the maturation of human dendritic cells. J Immunol
167: 4844–4852, 2001.
72. Srivastava PK. Role of heat-shock proteins in innate and adaptive
immunity. Nat Rev Immunol 2: 185–194, 2002.
73. Srivastava PK, DeLeo AB, and Old LJ. Tumor rejection antigens of
chemically induced tumors of inbred mice. Proc Natl Acad Sci USA 83:
3407–3411, 1986.
74. Srivastava PK, Menoret A, Basu S, Binder RJ, and McQuade KL.
Heat shock proteins come of age: primitive functions acquire new roles in
an adaptive world. Immunity 8: 657–665, 1998.
75. Srivastava PK, Udono H, Blachere NE, and Li Z. Heat shock proteins
transfer peptides during antigen processing and CTL priming. Immunogenetics 39: 93–98, 1994.
76. Stelter F. Structure/function relationships of CD14. Chem Immunol 74:
25–41, 2000.
77. Szewezuk MR and Depew WT. Evidence for T lymphocyte reactivity to
the 65 kilodalton heat-shock protein of mycobacteria in active Crohn’s
disease. Clin Invest Med 15: 494–505, 1992.
78. Tapping RI, Akashi S, Miyake K, Godowski PJ, and Tobias PS.
Toll-like receptor 4, but not Toll-like receptor 2, is a signaling receptor for
Escherichia and Salmonella lipopolysaccharide. J Immunol 165: 5780–
5787, 2000.
79. Triantafilou K, Triantafilou M, and Dedrick RL. A CD14-independent
LPS receptor cluster. Nat Immunol 2: 338–345, 2001.