ArticlePDF Available

Pivotal Advance: Analysis of proinflammatory activity of highly purified eukaryotic recombinant HMGB1 (amphoterin)

Authors:
  • Omnia, The Joint Authority of Education in Espoo Region

Abstract and Figures

HMGB1 (amphoterin) is a 30-kDa heparin-binding protein that mediates transendothelial migration of monocytes and has proinflammatory cytokine-like activities. In this study, we have investigated proinflammatory activities of both highly purified eukaryotic HMGB1 and bacterially produced recombinant HMGB1 proteins. Mass analyses revealed that recombinant eukaryotic HMGB1 has an intrachain disulphide bond. In mass analysis of tissue-derived HMGB1, two forms were detected: the carboxyl terminal glutamic acid residue lacking form and a full-length form. Cell culture studies indicated that both eukaryotic and bacterial HMGB1 proteins induce TNF-alpha secretion and nitric oxide release from mononuclear cells. Affinity chromatography analysis revealed that HMGB1 binds tightly to proinflammatory bacterial substances. A soluble proinflammatory substance was separated from the bacterial recombinant HMGB1 by chloroform-methanol treatment. HMGB1 interacted with phosphatidylserine in both solid-phase binding and cell culture assays, suggesting that HMGB1 may regulate phosphatidylserine-dependent immune reactions. In conclusion, HMGB1 polypeptide has a weak proinflammatory activity by itself, and it binds to bacterial substances, including lipids, that may strengthen its effects.
HMGB1-binding substances and cytokine expression. HMGB1 binds to bacterial proinflammatory substances. E. coli homogenates were applied to HMGB1 or sepharose CL-4B columns, the columns were washed with 0.5 M NaCl-TBS (0.5 M wash), and bound substances were eluted with increasing salt concentrations. TNF-induction by coated wash and elution fractions was determined in macrophage cell culture assay (A). HMGB1 column fractions black bars; Sepharose CL-4B column fractions are denoted by open bars. An active substance can be extracted by chloroform-methanol-water partition to polar lipid phase from bacterially produced HMGB1. Five migrograms of recHMGB1 or deltaC-HMGB1 purified using glutathione-sepharose chromatography (in 100 l of PBS) was treated with chloroform-methanol mixture, and water was added to separate nonpolar and polar phases. The polar phase was dried to plastic tubes. Induction of macrophage TNF-secretion was assayed using RAW 264.7 cells, and secreted TNF-was quantified using ELISA (B). Samples from control partitions were used as controls, and their TNF-release was determined as 100%. (n3; *, P0.05 when compared with control). HMGB1 binds to acidic phospholipids. Binding of HMGB1 to phospholipid-coated wells was detected by ELISA (C). Phospholipids were dissolved in methanol, and various amounts of lipids (0, 100, or 300 g) were dried on microwells. The wells were blocked with BSA and incubated with 2 g/ml of recHMGB1 for 1h. Bound recHMGB1 was detected with antipeptide I (squares) and antipeptide III (triangles) ELISAs. Results of control ELISA without primary antibody are indicated by solid circles. PA, phosphatitic acid; PE, phosphatidylethanolamine; PS, phosphatidylserine ; n3. *, P0.05 when compared with control wells. Effect of HMGB1 on phosphatidylserine-mediated inhibition of LPS-induced macrophage nitric oxide release (D). LPS (10 ng/ml)-activated RAW 264.7 cells were cultured in the presence of 30, 60, or 120 g/ml of PS-containing vesicles, or in the presence of 60 or 120 g/ml of phosphatidylcholine (PC) vesicles. Nitric oxide in culture medium was measured after 20 h, and the results were normalized to values of cell cultures without added lipids. PS inhibited nitric oxide release dose dependently when compared with PC control (solid bars). Effect of HMGB1 (30 g/ml) on the PS-mediated inhibition was tested (open bars) (n3).
… 
Content may be subject to copyright.
Pivotal Advance: Analysis of proinflammatory activity of highly
purified eukaryotic recombinant HMGB1 (amphoterin)
Ari Rouhiainen,
1
Sarka Tumova, Leena Valmu, Nisse Kalkkinen, and Heikki Rauvala
Neuroscience Center, and Institute of Biotechnology, University of Helsinki, Helsinki, Finland
Abstract: HMGB1 (amphoterin) is a 30-kDa hepa-
rin-binding protein that mediates transendothelial
migration of monocytes and has proinflammatory cy-
tokine-like activities. In this study, we have investi-
gated proinflammatory activities of both highly puri-
fied eukaryotic HMGB1 and bacterially produced
recombinant HMGB1 proteins. Mass analyses re-
vealed that recombinant eukaryotic HMGB1 has an
intrachain disulphide bond. In mass analysis of tissue-
derived HMGB1, two forms were detected: the car-
boxyl terminal glutamic acid residue lacking form
and a full-length form. Cell culture studies indicated
that both eukaryotic and bacterial HMGB1 proteins
induce TNF-secretion and nitric oxide release from
mononuclear cells. Affinity chromatography analysis
revealed that HMGB1 binds tightly to proinflamma-
tory bacterial substances. A soluble proinflammatory
substance was separated from the bacterial recombi-
nant HMGB1 by chloroform-methanol treatment.
HMGB1 interacted with phosphatidylserine in both
solid-phase binding and cell culture assays, suggesting
that HMGB1 may regulate phosphatidylserine-
dependent immune reactions. In conclusion,
HMGB1 polypeptide has a weak proinflammatory
activity by itself, and it binds to bacterial sub-
stances, including lipids, that may strengthen its
effects. J. Leukoc. Biol. 81: 49 –58; 2007.
Key Words: inflammation nitric oxide phosphatidylserine
RAGE TNF-
INTRODUCTION
HMGB1 (amphoterin) is a 30-kDa heparin-binding protein
widely expressed in different tissues and organisms [1–3].
Macrophages, monocytes, endothelial cells, neurons, and var-
ious tumor cells secrete HMGB1 after induction, and high
amounts of HMGB1 have been detected in serum samples of
various inflammatory diseases [4 –10].
Recent studies have highlighted the role of HMGB1 as a
monocyte and endothelium-activating substance, and mediator
of inflammation [4, 11, 12, and as reviewed in 13 and 14].
HMGB1 induces inflammation during necrosis, whereas in
apoptotic cells, it is sequestered to the nucleus [11]. During
trauma and inflammation, HMGB1 is massively released to
extracellular space, where it mediates organ damage and le-
thality [4, 15]. Further, elevated levels of extracellular HMGB1
or HMGB1 mRNA are detected in cancer and in inflammatory
disorders, suggesting that HMGB1 is acting as a widespread
inflammatory mediator [16, 17].
HMGB1 binds to several transmembrane receptors, includ-
ing the receptor for advanced glycation end products (RAGE),
Toll-like receptors 2 and 4 (TLR2/4), and syndecan-1 (CD138),
and generates proinflammatory signaling to nucleus [18 –22].
Well-characterized signaling routes of HMGB1 are NF-B and
ERK1/2 activation by RAGE-ligation, and IKKa/b and NF-B
activation by TLRs [19, 23–25]. In addition, HMGB1 and
RAGE mediate transendothelial migration of monocytes and
tumor cells [5, 26].
HMGB1 has a characteristic bipolar structure (amphoterin),
and it avidly binds to various substances like DNA or heparin
[2]. Further, it is often post-translationally modified [27].
Whether HMGB1 in solid tissue or circulation binds other
inflammation mediators is currently unclear. HMGB1 derived
from eukaryotic cells is less proinflammatory than the recom-
binant protein derived from bacterial expression systems [28,
29]. This suggests that the different forms of HMGB1 or the
existence of HMGB1 binding cofactors influence proinflamma-
tory activity.
We have produced recombinant HMGB1 (recHMGB1) in a
baculovirus system yielding high expression levels (50 –100
mg/l) of recombinant protein [10]. In addition, we have pro-
duced an E. coli-derived recombinant, and purified HMGB1
from tissue. In this study, we have tested the effect of both
tissue derived and recombinant HMGB1 proteins in mononu-
clear cell proinflammatory responses.
MATERIALS AND METHODS
Materials
ATP (ATP), lipopolysaccharide (LPS), phosphatidic acid, phosphatidylethanol-
amine, polymyxin B and S100b were from Sigma-Aldrich (St. Louis, MO,
USA). Phosphatidylserine was from Avanti Polar Lipids (Alabaster, AL, USA).
Advanced glycation end bovine serum albumin (AGE-BSA) was produced as
described [24]. Recombinant HMGB1 (recHMGB1) was produced and purified
as described, and analyzed in GelCode Blue (Pierce, Rockford, IL)-stained
SDS-PAGE [5, 10]. cDNA coding for amino terminal amino acids 1-185 of
HMGB1 was cloned into pGEX-6P-1-plasmid. Recombinant GST-fusion pro-
1
Correspondence: Neuroscience Center, Viikinkaari 4, PL 56, University of
Helsinki, Helsinki 00014, Finland. E-mail: ari.rouhiainen@helsinki.fi
Received March 14, 2006; revised August 9, 2006; accepted August 9,
2006.
doi: 10.1189/jlb.0306200
0741-5400/07/0081-0049 © Society for Leukocyte Biology Journal of Leukocyte Biology Volume 81, January 2007 49
tein (deltaC-HMGB1) was expressed in BL21(pLysS) bacteria, and purified
using glutathione-sepharose column and PerScission Protease method (GE
Healthcare Bio-Sciences AB, Uppsala, Sweden). In some experiments deltaC-
HMGB1 was further purified with HiTrap Heparin and HiTrap SP chromatog-
raphies (GE Healthcare Bio-Sciences AB, New York, NY). The truncated
HMGB1 (deltaC-HMGB1) is still capable to trigger RAGE signaling [26], but
it is more soluble than the nascent HMGB1. Tissue HMGB1 was isolated from
E18-P2 rat brain using heparin-sepharose and Affi-Gel Blue chromatographies
(GE Healthcare Bio-Sciences AB) as described, except that the NaDH washing
step was omitted [30]. DNA content in recombinant protein stocks were
measured using CyQuant Kit (Promega, Madison, WI). Both recHMGB1 and
deltaC-HMGB1 contained 0.03 g DNA per 1 mg of protein. Endotoxin
content in recHMGB1 fraction was under detection limit [5].
Cells and cell culture
RAW 264.7 cells were cultured as described [5]. Mixed rat glial cultures were
prepared from neonatal rat brain and cultured as described [31]. PBMCs were
isolated from adult male NMRI mice. Blood was collected; mononuclear cells
were enriched using a method described by Graziani-Bowering et al., and cells
were washed with 1 mM EDTA-PBS [32]. In some assays, mononuclear cell
fractions from two mice were pooled. Cells in 10% FCS-RPMI were adhered to
HMGB1-coated or control cell culture wells.
Primary and secondary structure analyses
HMGB1 used in mass analyses was analyzed and purified using RP-HPLC [10,
33]. In some studies recHMGB1 was reduced and alkylated. Alkylation was
done using 4-vinylpyridine alkylating agent, which causes 105 Da or 106 Da
increase in mass when bound to nonreduced or reduced cysteine residue,
respectively [34]. Trypsin digestion and mass spectrometric analyses were
carried out as described [33, 34]. To determine disulfide-bonded cysteines,
tryptic peptides derived from nonreduced recHMGB1 were analyzed in mass
spectrometry.
Database searches
Expressed sequence tag (EST) searches of nucleotide databases were done
using tools from the Web sites of National Center for Biotechnology Informa-
tion (Rockville Pike, Bethesda, MD).
Heparin-binding experiments
Heparin-sepharose chromatographies were done as described previously [30]
using A
¨kta chromatography station and 1 ml HiTrap heparin column (GE
Healthcare Bio-Sciences AB). In some experiments, samples contained 1 mM
dithiothreitol or 10 mM -mercaptoethanol.
TNF-induction and secretion assays
In HMGB1-induced RAW 264.7 macrophage secretion assays, 0.1– 0.2 10
6
cells (in OPTIMEM I medium, Invitrogen, Carlsbad, CA) were cultured in the
wells of cell culture plates. Proteins or LPS were added in the medium as
defined in each experiment, and cells were cultured for indicated times. In
some studies, reduced (3% -mercaptoethanol treated) proteins were coated to
wells for 1 h and washed before cell culture. TNF-concentration in culture
medium was measured using ELISA (Bender Medsystems, Vienna, Austria).
TNF-standards used in ELISA were from Bender Medsystems, Roche
(Mannheim, Germany) and Endogen (Pierce). Resulting TNF-values were
normalized to values from uninduced cells, which was defined as 100%.
Mouse PBMC TNF-mRNA induction and
protein secretion
Cells were adhered to HMGB1-coated or control 48-well plate wells (3–7
wells/mouse total PBMCs) in 10% FCS-RPMI and cultured for 6 h. TNF-in
culture media was quantified using ELISA. In some assays, mRNA coding for
TNF-was detected from PBMC-cultures by RT-PCR. Cells were cultured for
6 h, and RT-PCR analysis was done as described [5]. Primers for TNF-were
acagaaagcatgatccgcgacg and ggctcagccactccagctgctc. Amplified DNA was an-
alyzed in agarose gel electrophoresis, and relative OD values of bands were
measured as described [5]. Porphobilinogen deaminase housekeeping gene
was used as a control, and TNF-values were normalized to porphobilinogen
deaminase values. The OD value of uninduced cells was defined as 1.
Interleukin-6 (IL-6) and monocyte chemotactic
protein-1 (MCP-1) expression analyses
RAW 264.7 cells were cultured for 2 days in OPTI-MEM I with or without
coated recHMGB1 (20 g/ml), S100b (20 g/ml), AGE-BSA (500 g/ml) or
with soluble LPS (0.1 g/ml). Equal amounts of RNA were reverse transcribed
and analyzed in RT-PCR. The oligopairs for IL-6 and MCP-1 were cagttgcct-
tcttgggactgatgctg and agcatccatcatttctttgtatctctgg, and atgcaggtccctgtcatgct-
tctgg and ggtgctgaagaccttagggcagatg, respectively. IL-6 and MCP-1 values
were normalized to porphobilinogen deaminase values [5]. The OD value of
uninduced cells was defined as 1. Trimmean values, excluding 17% of lowest
and highest values, were calculated for IL-6 samples, and mean values were
calculated for MCP-1 samples.
Nitric oxide secretion assay
RAW 264.7 or mixed rat glial cultures were treated with various amounts of
HMGB1 proteins or LPS and cultured for indicated times. RAW 264.7 cells
were cultured in RPMI or DMEM supplemented with 10% FCS, and PBMCs
were cultured in RPMI supplemented with 10% FCS. 10 g/ml of polymyxin
B was added to some assays. Nitric oxide was quantified from culture media
using Greiss Reagent System (Promega). Inducible nitric oxide synthase
(iNOS) expression levels were analyzed with RT-PCR using the primers
atggcttgcccctggaagttt and ggcttgtctctgggtcctctggt. Amplified DNA was normal-
ized and quantified as described above.
Coculture of endotoxin-activated RAW 264.7
cells with phospholipid vesicles and recHMGB1
Cells were cultured on microwell plates in 10% FCS-RPMI (1.510
5
cells/
well). LPS (10 ng/ml) was added to all of the wells. Phospholipid vesicles
containing phosphatidylcholine alone or phosphatidylcholine (70%) and phos-
phatidylserine (30%) were made as described with two exceptions: the chlo-
roform was evaporated with nitrogen gas, and filter pore size used was 0.2 m
[35]. Various amounts of phospholipid vesicles with or without 30 g/ml of
recHMGB1 were added, and cells were cultured for 20 h. Nitric oxide in
culture media was quantified using Greiss Reagent System. Values from cell
cultures without added lipids were determined as 1, and values of lipid
containing wells were normalized to this value in both control and recHMGB1
samples.
Extraction of HMGB1-binding bacterial
substances
XL-1 E. Coli cells were homogenized in cold TBS containing lysozyme (Sigma)
and protease inhibitors [36]. RecHMGB1 was coupled to EAH-sepharose (GE
Healthcare Bio-Sciences AB) at the concentration of 1 mg per 1 ml of
sepharose gel. The cleared soluble fraction was applied to recHMGB1 column.
The column was washed with TBS containing 0.5 M NaCl, and bound sub-
stances were eluted with increasing salt concentrations. Uncoupled sepharose
CL-4B (Sigma) was used as a control column. DNA content in the fractions was
measured using CyQuant assay kit. For macrophage TNF-secretion induc-
tion measurement, diluted fractions were coated to plastic wells. Macrophages
were adhered to wells and cultured, and TNF- was measured by ELISA.
RecHMGB1 or glutathione-sepharose column and PerScission Protease
method purified deltaC-HMGB1 were used as samples in chloroform-methanol
extraction. Extraction was done using the method described by M. Pasciak et
al. with some modifications [37]. Briefly, 5 g of protein was diluted in 100 l
of PBS in polypropylene tube, and 600 l of chloroform-methanol (2:1) was
added with mixing. Then, 200 l of methanol and 300 l of water was added,
tubes were mixed by vortexing, and centrifuged at 16,000 gfor 5 min. The
upper phase was dried to plastic tubes, and 4 10
5
RAW 264.7 cells in 400
l of OPTIMEM I was added to tubes. Tubes were mixed after 30 min
incubation, and after 60 min incubation, cell suspensions were transferred to
a 48-well plate and subcultured for 2 h. TNF-was quantified from culture
supernatants using ELISA.
Phospholipid binding assay
The assay was carried out essentially as described by Nakano et al. [38].
Phospholipids were dissolved in methanol, and various amounts of lipids (0,
50 Journal of Leukocyte Biology Volume 81, January 2007 http://www.jleukbio.org
100, or 300 g) were added in 100 l to microwells and dried. Wells were
blocked with 1% BSA-PBS, and 2 g/ml of recHMGB1 in 1% BSA-PBS was
added to the wells for 1 h. The wells were washed, and bound recHMGB1 was
detected with antipeptide I and antipeptide III ELISAs [10]. Wells omitting the
primary antibody were used as negative controls.
Statistics
Pvalues were calculated using Student's unpaired ttest in Microsoft
Excel2000 program (Microsoft Corporation, Redmond, WA). Error bars in all
figures represent means SD.
RESULTS
Structural analysis of HMGB1
Recombinant HMGB1 proteins were produced in either S9
baculovirus or E. coli expression systems. The baculovirus
system produced extremely high levels of full-length recH-
MGB1, yielding 50 –100 mg/l of recombinant protein in culture
stocks from which it was purified with heparin-sepharose and
ion exchange chromatography [10]. Tissue-HMGB1 was iso-
lated from young rat brain with a two-step chromatography
method using heparin-sepharose and Affi Gel Blue chromatog-
raphy [30]. Both recHMGB1 and tissue-HMGB1 migrated as a
single band in SDS-PAGE both under nonreducing and reduc-
ing conditions (Fig. 1A and data not shown). The finding that
proteins migrated faster under reducing conditions suggests
that the oxidation state is the same in both eukaryotic proteins.
Mass spectrometric analyses of tissue-HMGB1 suggested that,
compared with the recHMGB1, the major form lacks the car-
boxyl terminal glutamic acid residue (the mass was 129 Da
lower than excepted), and the minor form is the full-length
protein. Results of tissue HMGB1 analyses are similar to
results by Chou et al. [39]. No ESTs coding for glutamic acid
residue-lacking form of rat HMGB1 were found in databases,
suggesting that lack of the glutamic acid residue is not due to
modifications of the HMGB1 transcript (data not shown).
Primary structure of recHMGB1 protein was analyzed using
RP-HPLC and mass spectrometry. RecHMGB1 was eluted at
two close peaks in RP-HPLC, which both had identical masses
of 24760 (Fig. 1, B and E, and data not shown). After reduction
and alkylation, recHMGB1 was eluted as a single peak (Fig.
1C), suggesting that different retention times of the native
protein fractions may be due to differences in protein confor-
mation. Tissue-HMGB1 was eluted at three peaks (Fig. 1D) as
described earlier [10]. The most prominent peak was identified
as HMGB1 (data not shown and Ref. 10).
The mass of the RP-HPLC purified recHMGB1 was 2–3 Da
lower than the calculated theoretical mass, suggesting one
intrachain disulphide bond (Fig. 1E). Mass-spectrometric anal-
yses of the trypsinized recHMGB1 peptides showed that the
disulfide bond exists between the first two cysteines within the
A-box in the HMGB1 derived from the first peak of RP-HPLC
(data not shown).
We tested whether reduction has any effect on heparin
binding. The reduced and nonreduced recHMGB1 bound to
heparin-sepharose with a similar affinity, indicating that hep-
arin binding of HMGB1 is redox state independent (Fig. 1F).
Effect of HMGB1 proteins on TNF-secretion
from mononuclear cells
The eukaryotic and bacterial recombinant HMGB1 proteins
and the tissue-derived protein were tested for their effect on
TNF-secretion using RAW 267.4 cells or mouse PBMCs.
The full-length bacterial recombinant was somewhat more ac-
tive than the deleted form (deltaC-HMGB1), but both recom-
binants rapidly induced TNF-secretion when added in solu-
tion or coated on the substrate (shown for deltaC-HMGB1 in a
3-h assay in Fig. 2A). In contrast, no significant TNF-
induction was observed under the same conditions for the
eukaryotic recombinant HMGB1 in 3-h assays (Fig. 2A). How-
ever, a TNF-inducing activity that was slightly above the
baseline was observed for the eukaryotic proteins in long-term
assays (shown for the recombinant and tissue-derived protein
in an 8-h assay in Fig. 2B). The reducing agent -mercapto-
ethanol had no effect on the ability of recombinant HMGB1
proteins to induce TNF-secretion (data not shown). An
inducing effect on TNF-secretion and mRNA level was also
observed for eukaryotic HMGB1 in PBMCs in 6-h assays (Fig.
2, C and D).
RecHMGB1 does not induce expression of IL-6
and MCP-1
RAGE ligation has been shown to induce expression of proin-
flammatory cytokines IL-6 and MCP-1 [40, 41]. We tested
using RT-PCR whether coated recHMGB1 is capable of in-
ducing IL-6 and MCP-1 in macrophages. AGE-BSA and S100b
were used as RAGE-ligand controls, and LPS was used as a
control for macrophage activation. Both AGE-BSA and LPS
induced significant upregulation of IL-6 and MCP-1 after 2
days of culture, whereas recHMGB1 and S100b had no effect
(Table 1). Interestingly, S100b and S100A1 were recently
shown to be incapable of inducing cytokine production [42, 43].
HMGB1 induces nitric oxide release from
macrophages
We tested whether recombinant HMGB1 proteins induce nitric
oxide release from both primary cells and transformed macro-
phages. A slow nitric oxide secretion was seen in RAW 264.7
cell cultures with high concentrations of soluble eukaryotic
HMGB1 (Fig. 3A), and it was not inhibited by polymyxin B
(data not shown). Bacterially produced soluble HMGB1 was a
more potent nitric oxide inducer (Fig. 3B). Effect of eukaryotic
soluble HMGB1 proteins on iNOS mRNA expression was
tested. RAW 264.7 cells were treated with 100 g/ml of
recHMGB1 or 10 or 100 ng/ml of LPS for 19 h, and iNOS
mRNA was quantified using RT-PCR analysis. Values ob-
tained were 104 1.3% (P0.05), 111 4.6%, and 115
5.5% [for 100 g/ml HMGB1, 10 ng/ml of LPS, and 100 ng/ml
of LPS, respectively (n3)] when uniduced expression was
determined as 100%. In mixed rat glial primary cell cultures,
soluble recHMGB1 induced nitrite release at lower concentra-
tions than from RAW 264.7 cells (Fig. 3C).
HMGB1 binds to bacterial proinflammatory
substances
Because the bacterially produced recombinant HMGB1 in-
duced strong proinflammatory reactions in mononuclear cells,
Rouhiainen et al. HMGB1 and cytokine expression 51
Fig. 1. Structural characteristics of recombinant HMGB1. (A)
Analysis of recHMGB1 (3.25 micrograms) or tissue derived
HMGB1 (1 microgram) was performed in GelCode Blue stained
SDS-PAGE. Proteins migrate as 30 kDa bands. (B–D) RP-HPLC
analysis of native HMGB1 proteins, and reduced and alkylated
recHMGB1. recHMGB1 elutes at two close peaks in RP-HPLC (B).
After reduction and alkylation of cysteine residues recHMGB1
elutes as a single peak (C). Tissue derived HMGB1 eluted as one
major peak. In addition, two minor peaks are detected (D). Time (t)
and absorbance (AU) axes in figures are not in scale. (E) The first
peak of recHMGB1 separated in RP-HPLC (B) was analyzed in
mass spectrometry. The data indicated the molecular mass of 24760 Da for recHMGB1. A second recHMGB1 peak from RP-HPLC
(B) gave an identical mass (data not shown). (F) Reduction does not influence recHMGB1 binding to heparin. Non-reduced or reduced
recHMGB1 was analyzed in heparin-Sepharose chromatography. The bound protein was eluted with 0.15–1.5 M NaCl gradient. All
recHMGB1 samples were eluted at the same NaCl concentration (0.7 M NaCl).
52 Journal of Leukocyte Biology Volume 81, January 2007 http://www.jleukbio.org
we tested whether HMGB1 is capable of binding macrophage-
activating bacterial substances. HMGB1 affinity chromatogra-
phy revealed that some bacterial components bind tightly to
HMGB1, and when released from HMGB1, they can elicit a
proinflammatory response (Fig. 4A). One such substance,
DNA, was detected from fractions eluted from the HMGB1
affinity column (data not shown).
Eukaryotic recHMGB1 and the bacterial recombinant pro-
teins purified with glutathione-sepharose column and PerScis-
sion Protease method were treated with chloroform-methanol to
denature the proteins and to separate possible polypeptide-
bound lipophilic substances. Both organic and polar phases
were separated with the addition of water. The polar phase from
the bacterial recombinant was found to induce macrophage
TNF-secretion. No such activity was found in eukaryotic
recHMGB1 or bacterial protein that was further purified with
heparin- and ion-exchange chromatography (Fig. 4B and data
not shown).
HMGB1 has been previously shown to bind both phospha-
tidylserine and sulfatide lipids [44, 45]. In this study, we tested
HMGB1 binding to three phospholipids expressed in E. coli:
phosphatidic acid, phosphatidylethanolamine, and phosphati-
dylserine [46]. HMGB1 bound strongly to phosphatidylserine
as excepted. In addition, HMGB1 bound strongly to phospha-
tidic acid. Binding to phosphatidylethanolamine was much
weaker (Fig. 4C).
Phosphatidylserine is a well-known immune suppressor
[47]. Therefore, we tested the effect of recHMGB1 on phos-
phatidylserine-mediated inhibition of nitric oxide release from
macrophages [48]. Phosphatidylserine vesicles inhibited LPS
induced nitric oxide release from RAW 264.7 cells dose de-
pendently (Fig. 4D). Coincubation with recHMGB1 affected
Fig. 2. HMGB1 and TNF-expression. Bacterially
produced recombinant HMGB1is a more potent TNF-
secretion inducer than eukaryotic recombinant
HMGB1. RAW 264.7 cells (1–2 10
5
cells/well in
OPTIMEM I) were added to HMGB1-coated (20 or 100
g/ml) microwells, and cultured for 3 h (A). TNF-
concentration in the culture medium was measured
using ELISA. Values from control wells were deter-
mined as 100%, and sample values were normalized
to control values. Solid bars denote 20 g/ml; open
bars denote 100 g/ml. (n3, *, P0.04 when com-
pared with controls, #, P0.05 when compared with
recHMGB1 samples). Eukaryotic recHMGB1 in-
duces TNF-release from macrophages. 20 g/ml of
recHMGB1 or tissue-HMGB1 was added to RAW
264.7 cultures (1–210
5
cells/well in OPTIMEM I).
LPS (0.1 g/ml) was added to positive control cultures.
Cells were cultured for 8 h (B). TNF-concentration in
the culture medium was measured using ELISA. Values
from wells without activators were determined as 100%,
and sample values were normalized to nonactivated
control values (n5; *, P0.01). Tissue-derived
HMGB1 induces TNF-secretion from mouse PBMCs.
Freshly isolated mouse PBMCs in 10% FCS-RPMI
were adhered to tissue-HMGB1-coated (20 g/ml)
plastic wells and cultured (C). After 6 h of culture,
TNF-concentration in the medium was measured.
Results were calculated as in Fig. 2B. (n4, *,
P0.03). HMGB1 induces TNF-mRNA in mouse
PBMCs. Cells were isolated and cultured as described
in Fig. 3C. RNA was isolated and analyzed in RT-PCR,
and relative OD values of the bands were measured (D).
OD value of uninduced cells was determined as 1. ODs
of bands were normalized to porphobilinogen mRNA
bands. *, P0.05 when compared with uninduced
sample. (n3).
TABLE 1. recHMGB1 or S100b Does Not Induce IL-6 or MCP-1 Gene Expression After Culture of 2 Days
a
HMGB1 S100b AGE-BSA LPS
IL-6 0.94 0.12 (Pns) 0.90 0.46 (Pns) 2.09 0.21 (P0.03) 7.70 2.29 (P0.03)
MCP-1 0.94 0.17 (Pns) 0.97 0.22 (Pns) 2.39 0.65 (P0.03) 2.45 0.57 (P0.03)
a
RAW 264.7 cells were cultured for 2 days in OPTI-MEM I with recHMGB1 (20 g/ml), S100b (20 g/ml), AGE-BSA (500 g/ml), or LPS (0.1 g/ml). Equal
amounts of RNA were analyzed in IL-6 and MCP-1 RT-PCR, and relative OD values of the bands were measured and normalized to values of the housekeeping
gene porphobilinogen deaminase. The OD value of uninduced samples was defined as 1.
Rouhiainen et al. HMGB1 and cytokine expression 53
only slightly the inhibitory effect of phosphatidylserine (Fig.
4D).
DISCUSSION
In the current study, we have taken advantage of the high
expression level of HMGB1 in our baculovirus expression
system in insect cells, which is expected to reduce the risk of
contaminating substances in the recombinant protein. The
expression system made it possible to isolate the recombinant
in a highly purified form using mild nondenaturing conditions,
without using trichloroacetic acid that is commonly used to
purify HMG-type proteins. For example, plasminogen activa-
tion-enhancing effect by HMGB1 and its DNA binding capa-
bility are strongly affected by acid treatment of the protein [10,
49]. Further, we have purified tissue-derived HMGB1 from rat
brain and show that it is very similar to recHMGB1.
Structural studies of the baculovirus-derived HMGB1 pro-
duced in animal cells show that it contains an intrachain
disulfide bond. Occurrence of the disulfide bond has been also
demonstrated in HMGB1 isolated from tissue [30, 50, 51]. The
molecular mass of the recombinant HMGB1 corresponds ex-
actly to the calculated molecular mass and does not give
evidence of other post-translational modifications, than the
presence of one disulfide bond. The tissue-derived HMGB1
used in this study differs from the recombinant eukaryotic
protein in that the major form lacks the carboxyl terminal
glutamic acid residue. The minor form is the full-length pro-
tein.
Post-translational modifications may regulate induction of
inflammatory reactions by HMGB1 [27]. The results of this
study indicate that genuine reduced and oxidized HMGB1
polypeptides are weak TNF-and nitric oxide-inducing agents
in mononuclear cells. Further studies are warranted to reveal
how post-translational modifications affect proinflammatory ac-
tivity of HMGB1.
Treatment with reducing agents has no effect on HMGB1’s
heparin binding activity or proinflammatory activity, suggest-
ing that HMGB1 can preserve its functions in the absence of
the disulfide bond. In addition, NMR studies have previously
shown that the recombinant A-box of HMGB1, having one
cysteine replaced with serine, folds in a manner that allows a
close contact of the cysteine with the replacing serine residue
[52]. This suggests that the disulfide formation is not necessary
for folding to such conformation where the two cysteines are in
close proximity. Further, CD spectroscopy studies revealed
that both the reduced and nonreduced recHMGB1 have essen-
tially the same -helical structure, suggesting that reduction
has no effect on secondary structure (Tumova and Rauvala,
unpublished results). In contrast, CD spectroscopy studies of
Fig. 3. RecHMGB1 induces nitric oxide release and up-
regulates iNOS in macrophage cultures. Time course study
of HMGB1-induced nitric oxide release. RAW 264.7 cells
were cultured in the presence of various amounts of soluble
recHMGB1 or LPS, and nitric oxide was quantified from cul-
ture media after 1, 2, or 3 days of culture (A). Control cells are
denoted by solid bars; 0.1 g/ml of recHMGB1 is denoted by
open bars; 1 g/ml of recHMGB1 is denoted by dark gray bars;
10 g/ml of recHMGB1 is denoted by light gray bars; 100
g/ml of recHMGB1 is denoted by checkered bars; and 1
g/ml of LPS is denoted by ruled bars (n3; *, P0.05 when
compared with control samples). Bacterially produced HMGB1
is a potent nitric oxide inducer. RAW 264.7 cell were cultured
in the presence of soluble deltaC-HMGB1 (20 g/ml) or LPS
(0.1 g/ml) for 24 h (B). Nitric oxide in culture media was
quantified. (n4; *, P0.005 when compared with control
samples). RecHMGB1 induces nitric oxide release from pri-
mary cell cultures. Mixed glial cell cultures from neonatal rat
brains were incubated with LPS (1 g/ml) or recHMGB1 (1–30
g/ml) for 1 day, and nitric oxide was quantified from the
culture medium (C). The number of experiments is at least six
in all conditions tested. *, P0.05 when compared with sam-
ples without recHMGB1 or LPS.
54 Journal of Leukocyte Biology Volume 81, January 2007 http://www.jleukbio.org
perchloric acid-treated HMGB1 revealed major changes in
spectra after reducing of the disulfide bond [53].
Our current results and previous results from other groups
suggest that eukaryotic and bacterial HMGB1 proteins differ in
their ability to induce TNF-[28, 29]. Further, our results
show that a TNF--inducing activity can be extracted by a
lipid solvent (chloroform/methanol) from the bacterially ex-
pressed recombinant but not from the highly purified baculo-
virus-derived protein expressed in animal cells. It seems prob-
able that the extremely high expression level achieved in our
eukaryotic system largely overrides the occurrence of copuri-
fying factor(s) that enhance proinflammatory activity.
The occurrence of the proinflammatory activity in the polar
phase in Folch partition after chloroform-methanol extraction
of the bacterially produced recombinant suggests that the
activity is enhanced by a polar lipid. Tentative analysis of
lipids in this fraction using mass spectrometry reveals a com-
plex lipid pattern (A. Rouhiainen, H. Rauvala, and P. Somer-
harju, unpublished observations), and further work is war-
ranted to elucidate the molecular nature of the active compo-
Fig. 4. HMGB1-binding substances and cytokine expression. HMGB1 binds to bacterial proinflamma-
tory substances. E. coli homogenates were applied to HMGB1 or sepharose CL-4B columns, the columns
were washed with 0.5 M NaCl-TBS (0.5 M wash), and bound substances were eluted with increasing salt
concentrations. TNF-induction by coated wash and elution fractions was determined in macrophage cell
culture assay (A). HMGB1 column fractions black bars; Sepharose CL-4B column fractions are denoted
by open bars. An active substance can be extracted by chloroform-methanol-water partition to polar lipid
phase from bacterially produced HMGB1. Five migrograms of recHMGB1 or deltaC-HMGB1 purified
using glutathione-sepharose chromatography (in 100 l of PBS) was treated with chloroform-methanol
mixture, and water was added to separate nonpolar and polar phases. The polar phase was dried to plastic
tubes. Induction of macrophage TNF-secretion was assayed using RAW 264.7 cells, and secreted
TNF-was quantified using ELISA (B). Samples from control partitions were used as controls, and their TNF-release was determined as 100%. (n3;
*, P0.05 when compared with control). HMGB1 binds to acidic phospholipids. Binding of HMGB1 to phospholipid-coated wells was detected by ELISA
(C). Phospholipids were dissolved in methanol, and various amounts of lipids (0, 100, or 300 g) were dried on microwells. The wells were blocked with
BSA and incubated with 2 g/ml of recHMGB1 for 1h. Bound recHMGB1 was detected with antipeptide I (squares) and antipeptide III (triangles) ELISAs.
Results of control ELISA without primary antibody are indicated by solid circles. PA, phosphatitic acid; PE, phosphatidylethanolamine; PS, phospha-
tidylserine; n3. *, P0.05 when compared with control wells. Effect of HMGB1 on phosphatidylserine-mediated inhibition of LPS-induced macrophage
nitric oxide release (D). LPS (10 ng/ml)-activated RAW 264.7 cells were cultured in the presence of 30, 60, or 120 g/ml of PS-containing vesicles, or
in the presence of 60 or 120 g/ml of phosphatidylcholine (PC) vesicles. Nitric oxide in culture medium was measured after 20 h, and the results were
normalized to values of cell cultures without added lipids. PS inhibited nitric oxide release dose dependently when compared with PC control (solid bars).
Effect of HMGB1 (30 g/ml) on the PS-mediated inhibition was tested (open bars) (n3).
Rouhiainen et al. HMGB1 and cytokine expression 55
nents(s). Furthermore, we show that HMGB1 binds to purified
lipids in a microwell binding assay and interacts with phos-
phatidylserine in cell culture assay, agreeing with our previous
finding of HMGB1 binding to platelet lipids [44]. It appears
clear that HMGB1 binds at least phosphatidic acid and phos-
phatidylserine. Interestingly, phosphatidylserine has been im-
plicated in the regulation of inflammation [54, 55], and
HMGB1 might thus affect interactions of phosphatidylserine
with cells and regulate its anti-inflammatory activities.
HMGB1 is present in circulation during different inflamma-
tory diseases, but its function there is not fully understood [4,
6, 7]. Binding of HMGB1 to substances derived from microbes
and/or injured tissues might create complexes up-regulating
innate immune responses that, in turn, jeopardize tissue integ-
rity through production of toxic inflammatory mediators. For-
eign material binding capacity phenomenon for some circulat-
ing proteins, such as the LPS binding protein, vitronectin, and
fibronectin, is known to occur. These proteins have been shown
to strengthen macrophage responses to bacterial substances
[56, 57]. Furthermore, other highly charged recombinant pro-
teins, for example, heat shock proteins produced in bacterial
expression systems [58 61], have been shown to induce im-
mune cell activations through binding to microbe-derived sub-
stances.
Our nitric oxide induction results are similar to those of
Kuniyasu et al. who detected nitric oxide induction in macro-
phages by eukaryotic HMGB1 [23]. These results suggest that
nitric oxide synthase upregulation and nitric oxide release are
induced by high concentrations of HMGB1. Nitric oxide syn-
thase gene expression is regulated by NF-B [reviewed in 62].
Since the HMGB1 receptors TLR2/4 and RAGE activate NF-
B, it seems reasonable to assume that these receptors are
involved in HMGB1/macrophage signaling [19, 23, 63]. In fact,
Kuniyasu et al. have shown that HMGB1 activates NF-Bin
human monocytes [23]. Sumi and Ignarro have shown that
AGE-BSA-induced nitric oxide synthase upregulation is inhib-
ited by anti-RAGE antibodies in RAW 264.7 cells [64]. How-
ever, Adami et al. described that the RAGE ligand S100b-
induced nitric oxide release from microglial cells is RAGE
signaling independent but RAGE-ectodomain mediated, sug-
gesting that other cell surface receptors exist for S100b [65].
The recent study by Kim et al. showed that down-regulation of
HMGB1 by short hairpin RNA in postischemic brain decreases
iNOS mRNA expression, suggesting that HMGB1 is involved
in iNOS regulation in vivo [66].
ACKNOWLEDGMENTS
A. R. was supported by grants from the Aarne and Aili Tu-
runens´ Foundation and the Maud Kuistila Memorial Founda-
tion. H. R. was supported by grants from the Academy of
Finland, Finnish Cancer Organizations and the Sigrid Juse´lius
Foundation. We thank Seija Lehto and Eeva-Liisa Saarikalle
for excellent technical assistance.
REFERENCES
1. Bustin, M. (2002) At the crossroads of necrosis and apoptosis: signaling to
multiple cellular targets by HMGB1. Sci. STKE 151, PE39.
2. Merenmies, J., Pihlaskari, R., Laitinen, J., Wartiovaara, J., Rauvala, H.
(1991) 30-kDa heparin-binding protein of brain (amphoterin) involved in
neurite outgrowth. Amino acid sequence and localization in the filopodia
of the advancing plasma membrane. J. Biol. Chem. 266, 16722–16729.
3. Bianchi, M. E., Beltrame, M., Paonessa, G. (1989) Specific recognition of
cruciform DNA by nuclear protein HMG1. Science 243, 1056 –1059.
4. Wang, H., Bloom, O., Zhang, M., Vishnubhakat, J. M., Ombrellino, M.,
Che, J., Frazier, A., Yang, H., Ivanova, S., Borovikova, L., et al. (1999)
HMG-1 as a late mediator of endotoxin lethality in mice. Science 285,
248 –251.
5. Rouhiainen, A., Kuja-Panula, J., Wilkman, E., Pakkanen, J., Stenfors, J.,
Tuominen, R. K., Lepantalo, M., Carpen, O., Parkkinen, J., Rauvala, H.
(2004) Regulation of monocyte migration by amphoterin (HMGB1). Blood
104, 1174 –1182.
6. Sunden-Cullberg, J., Norrby-Teglund, A., Rouhiainen, A., Rauvala, H.,
Herman, G., Tracey, K. J., Lee, M. L., Andersson, J., Tokics, L., Treutiger,
C. J. (2005) Persistent elevation of high-mobility group box-1 protein
(HMGB1) in patients with severe sepsis and septic shock. Crit. Care Med.
33, 564 –573.
7. Yamada, S., Inoue, K., Yakabe, K., Imaizumi, H., Maruyama, I. (2003)
High-mobility group protein 1 (HMGB1) quantified by ELISA with a
monoclonal antibody that does not cross-react with HMGB2. Clin. Chem.
49, 1535–1537.
8. Mullins, G. E., Sunden-Cullberg, J., Johansson, A. S., Rouhiainen, A.,
Erlandsson-Harris, H., Yang, H., Tracey, K. J., Rauvala, H., Palmblad, J.,
Andersson, J., et al. (2004) Activation of human umbilical vein endothelial
cells leads to relocation and release of high-mobility group box chromo-
somal protein 1. Scand. J. Immunol. 60, 566 –573.
9. Rauvala, H., Merenmies, J., Pihlaskari, R., Korkolainen, M., Huhtala,
M. L., Panula, P. (1988) The adhesive and neurite-promoting molecule
p30: analysis of the amino-terminal sequence and production of antipep-
tide antibodies that detect p30 at the surface of neuroblastoma cells and
of brain neurons. J. Cell Biol. 107, 2293–2305.
10. Parkkinen, J., Raulo, E., Merenmies, J., Nolo, R., Kajander, E. O.,
Baumann, M., Rauvala, H. (1993) Amphoterin, the 30-kDa protein in a
family of HMG1-type polypeptides. Enhanced expression in transformed
cells, leading edge localization, and interactions with plasminogen acti-
vation. J. Biol. Chem. 268, 19726 –19738.
11. Scaffidi, P., Misteli, T., Bianchi, M. E. (2002) Release of chromatin protein
HMGB1 by necrotic cells triggers inflammation. Nature 418, 191–195.
12. Treutiger, C. J., Mullins, G. E., Johansson, A. S., Rouhiainen, A., Rauvala,
H. M., Erlandsson-Harris, H., Andersson, U., Yang, H., Tracey, K. J.,
Andersson, J., et al. (2003) High mobility group 1 B-box mediates acti-
vation of human endothelium. J. Intern. Med. 254, 375–385.
13. Lotze, M. T., Tracey, K. J. (2005) High-mobility group box 1 protein
(HMGB1): nuclear weapon in the immune arsenal. Nat. Rev. Immunol. 5,
331–342.
14. Sunden-Cullberg, J., Norrby-Teglund, A., Treutiger, C. J. (2006) The role
of high mobility group box-1 protein in severe sepsis. Curr. Opin. Infect.
Dis. 19, 231–236.
15. Tsung, A., Sahai, R., Tanaka, H., Nakao, A., Fink, M. P., Lotze, M. T.,
Yang, H., Li, J., Tracey, K. J., Geller, D. A., et al. (2005) The nuclear
factor HMGB1 mediates hepatic injury after murine liver ischemia-reper-
fusion. J. Exp. Med. 201, 1135–1143.
16. Hatada, T., Wada, H., Nobori, T., Okabayashi, K., Maruyama, K., Abe, Y.,
Uemoto, S., Yamada, S., Maruyama, I. (2005) Plasma concentrations and
importance of high mobility group box protein in the prognosis of organ
failure in patients with disseminated intravascular coagulation. Thromb.
Haemost. 94, 975–979.
17. Pachot, A., Monneret, G., Voirin, N., Leissner, P., Venet, F., Bohe, J.,
Payen, D., Bienvenu, J., Mougin, B., Lepape, A. (2005) Longitudinal study
of cytokine and immune transcription factor mRNA expression in septic
shock. Clin. Immunol. 114, 61– 69.
18. Hori, O., Brett, J., Slattery, T., Cao, R., Zhang, J., Chen, J. X., Nagashima,
M., Lundh, E. R., Vijay, S., Nitecki, D., et al. (1995) The receptor for
advanced glycation end products (RAGE) is a cellular binding site for
amphoterin. Mediation of neurite outgrowth and co-expression of rage and
amphoterin in the developing nervous system. J. Biol. Chem. 270,
25752–25761.
19. Park, J. S., Svetkauskaite, D., He, Q., Kim, J. Y., Strassheim, D., Ishizaka,
A., Abraham, E. (2004) Involvement of toll-like receptors 2 and 4 in
cellular activation by high mobility group box 1 protein. J. Biol. Chem.
279, 7370 –7377.
20. Park, J. S., Gamboni-Robertson, F., He, Q., Svetkauskaite, D., Kim, J. Y.,
Strassheim, D., Sohn, J. W., Yamada, S., Maruyama, I., Banerjee, A., et al.
(2006) High mobility group box 1 protein interacts with multiple Toll-like
receptors. Am. J. Physiol. Cell Physiol. 290, C917–C924.
56 Journal of Leukocyte Biology Volume 81, January 2007 http://www.jleukbio.org
21. Izuishi, K., Tsung, A., Jeyabalan, G., Critchlow, N. D., Li, J., Tracey, K. J.,
Demarco, R. A., Lotze, M. T., Fink, M. P., Geller, D. A., et al. (2006)
High-mobility group box 1 preconditioning protects against liver isch-
emia-reperfusion injury. J. Immunol. 176, 7154 –7158.
22. Salmivirta, M., Rauvala, H., Elenius, K., Jalkanen, M. (1992) Neurite
growth-promoting protein (amphoterin, p30) binds syndecan. Exp. Cell
Res. 200, 444 451.
23. Kuniyasu, H., Yano, S., Sasaki, T., Sasahira, T., Sone, S., Ohmori, H.
(2005) Colon cancer cell-derived high mobility group 1/amphoterin in-
duces growth inhibition and apoptosis in macrophages. Am. J. Pathol.
166, 751–760.
24. Huttunen, H. J., Fages, C., Rauvala, H. (1999) Receptor for advanced
glycation end products (RAGE)-mediated neurite outgrowth and acti-
vation of NF-kappaB require the cytoplasmic domain of the receptor
but different downstream signaling pathways. J. Biol. Chem. 274,
19919 –19924.
25. Ishihara, K., Tsutsumi, K., Kawane, S., Nakajima, M., Kasaoka, T. (2003)
The receptor for advanced glycation end-products (RAGE) directly binds
to ERK by a D-domain-like docking site. FEBS Lett. 550, 107–113.
26. Huttunen, H. J., Fages, C., Kuja-Panula, J., Ridley, A. J., Rauvala, H.
(2002) Receptor for advanced glycation end products-binding COOH-
terminal motif of amphoterin inhibits invasive migration and metastasis.
Cancer Res. 62, 4805– 4811.
27. Ulloa, L., Messmer, D. (2006) High-mobility group box 1 (HMGB1)
protein: Friend and foe. Cytokine Growth Factor Rev. 17, 189 –201.
28. Zimmermann, K., Volkel, D., Pable, S., Lindner, T., Kramberger, F.,
Bahrami, S., Scheiflinger, F. (2004) Native versus recombinant high-
mobility group B1 proteins: functional activity in vitro. Inflammation 28,
221–229.
29. Li, J., Wang, H., Mason, J. M., Levine, J., Yu, M., Ulloa, L., Czura, C. J.,
Tracey, K. J., Yang, H. (2004) Recombinant HMGB1 with cytokine-
stimulating activity. J. Immunol. Methods 289, 211–223.
30. Rauvala, H., Pihlaskari, R. (1987) Isolation and some characteristics of an
adhesive factor of brain that enhances neurite outgrowth in central neu-
rons. J. Biol. Chem. 262, 16625–16635.
31. Golde, S., Chandran, S., Brown, G. C., Compston, A. (2002) Different
pathways for iNOS-mediated toxicity in vitro dependent on neuronal
maturation and NMDA receptor expression. J. Neurochem. 82, 269 –282.
32. Graziani-Bowering, G. M., Graham, J. M., Filion, L. G. (1997) A quick,
easy and inexpensive method for the isolation of human peripheral blood
monocytes. J. Immunol. Methods 207, 157–168.
33. Ylonen, A., Rinne, A., Herttuainen, J., Bogwald, J., Jarvinen, M., Kalk-
kinen, N. (1999) Atlantic salmon (Salmo salar L.) skin contains a novel
kininogen and another cysteine proteinase inhibitor. Eur. J. Biochem.
266, 1066 –1072.
34. Ylonen, A., Kalkkinen, N., Saarinen, J., Bogwald, J., Helin, J. (2001)
Glycosylation analysis of two cysteine proteinase inhibitors from Atlantic
salmon skin: di-O-acetylated sialic acids are the major sialic acid species
on N-glycans. Glycobiology 11, 523–531.
35. Chen, X., Doffek, K., Sugg, S. L., Shilyansky, J. (2004) Phosphatidylserine
regulates the maturation of human dendritic cells. J. Immunol. 173,
2985–2994.
36. Froman, G., Switalski, L. M., Speziale, P., Hook, M. (1987) Isolation and
characterization of a fibronectin receptor from Staphylococcus aureus.
J. Biol. Chem. 262, 6564 6571.
37. Pasciak, M., Holst, O., Lindner, B., Mordarska, H., Gamian, A. (2003)
Novel bacterial polar lipids containing ether-linked alkyl chains, the
structures and biological properties of the four major glycolipids from
Propionibacterium propionicum PCM 2431 (ATCC 14157
T
).J. Biol. Chem.
278, 3948 –3956.
38. Nakano, T., Ishimoto, Y., Kishino, J., Umeda, M., Inoue, K., Nagata, K.,
Ohashi, K., Mizuno, K., Arita, H. (1997) Cell adhesion to phosphatidyl-
serine mediated by a product of growth arrest-specific gene 6. J. Biol.
Chem. 272, 29411–29414.
39. Chou, D. K., Evans, J. E., Jungalwala, F. B. (2001) Identity of nuclear
high-mobility-group protein, HMG-1, and sulfoglucuronyl carbohydrate-
binding protein, SBP-1, in brain. J. Neurochem. 77, 120 –131.
40. Schmidt, A. M., Hasu, M., Popov, D., Zhang, J. H., Chen, J., Yan, S. D.,
Brett, J., Cao, R., Kuwabara, K., Costache, G., et al. (1994) Receptor for
advanced glycation end products (AGEs) has a central role in vessel wall
interactions and gene activation in response to circulating AGE proteins.
Proc. Natl. Acad. Sci. USA 91, 8807– 8811.
41. Hou, F. F., Jiang, J. P., Guo, J. Q., Wang, G. B., Zhang, X., Stern,
D. M., Schmidt, A. M., Owen, W. F., Jr. (2002) Receptor for advanced
glycation end products on human synovial fibroblasts: role in the
pathogenesis of dialysis-related amyloidosis. J. Am. Soc. Nephrol. 13,
1296 –1306.
42. Valencia, J. V., Mone, M., Koehne, C., Rediske, J., Hughes, T. E. (2004)
Binding of receptor for advanced glycation end products (RAGE) ligands
is not sufficient to induce inflammatory signals: lack of activity of endo-
toxin-free albumin-derived advanced glycation end products. Diabetologia
47, 844 852.
43. Ehlermann P, Eggers K, Bierhaus A, Most P, Weichenhan D, Greten J,
Nawroth PP, Katus HA, Remppis A. (2006) Increased proinflammatory
endothelial response to S100A8/A9 after preactivation through advanced
glycation end products. Cardiovasc. Diabetol. 5, 6.
44. Rouhiainen, A., Imai, S., Rauvala, H., Parkkinen, J. (2000) Occurrence of
amphoterin (HMG1) as an endogenous protein of human platelets that is
exported to the cell surface upon platelet activation. Thromb. Haemost.
84, 1087–1094.
45. Mohan, P. S., Laitinen, J., Merenmies, J., Rauvala, H., Jungalwala, F. B.
(1992) Sulfoglycolipids bind to adhesive protein amphoterin (P30) in the
nervous system. Biochem. Biophys. Res. Commun. 182, 689 696.
46. Raetz, C. R., Dowhan, W. (1990) Biosynthesis and function of phospho-
lipids in Escherichia coli. J. Biol. Chem. 265, 1235–1238.
47. Savill, J., Dransfield, I., Gregory, C., Haslett, C. (2002) A blast from the
past: clearance of apoptotic cells regulates immune responses. Nat. Rev.
Immunol. 2, 965–975.
48. De, S. R., Ajmone-Cat, M. A., Nicolini, A., Minghetti, L. (2002) Expres-
sion of phosphatidylserine receptor and down-regulation of pro-inflamma-
tory molecule production by its natural ligand in rat microglial cultures.
J. Neuropathol. Exp. Neurol. 61, 237–244.
49. Wagner, J. P., Quill, D. M., Pettijohn, D. E. (1995) Increased DNA-
bending activity and higher-affinity DNA binding of high-mobility group
protein HMG-1 prepared without acids. J. Biol. Chem. 270, 7394 –7398.
50. Baker, C., Isenberg, I., Goodwin, G. H., Johns, E. W. (1976) Physical
studies of the nonhistone chromosomal proteins HMG-1 and HMG-2.
Biochemistry 15, 1645–1649.
51. Elton, T. S., Reeves, R. (1985) The effects of oxidation on the reverse-
phase high-performance liquid chromatography characteristics of the high
mobility groups 1 and 2 proteins. Anal. Biochem. 149, 316 –321.
52. Hardman, C. H., Broadhurst, R. W., Raine, A. R., Grasser, K. D., Thomas,
J. O., Laue, E. D. (1995) Structure of the A-domain of HMG1 and its
interaction with DNA as studied by heteronuclear three- and four-dimen-
sional NMR spectroscopy. Biochemistry 34, 16596 –16607.
53. Kohlstaedt, L. A., Sung, E. C., Fujishige, A., Cole, R. D. (1987) Non-
histone chromosomal protein HMG1 modulates the histone H1-induced
condensation of DNA. J. Biol. Chem. 262, 524 –526.
54. Gaipl, U. S., Beyer, T. D., Baumann, I., Voll, R. E., Stach, C. M., Heyder,
P., Kalden, J. R., Manfredi, A., Herrmann, M. (2003) Exposure of anionic
phospholipids serves as anti-inflammatory and immunosuppressive sig-
nal–implications for antiphospholipid syndrome and systemic lupus ery-
thematosus. Immunobiology 207, 73– 81.
55. Brouckaert, G., Kalai, M., Krysko, D. V., Saelens, X., Vercammen, D.,
Ndlovu, M., Haegeman, G., D'Herde, K., Vandenabeele, P. (2004) Phago-
cytosis of necrotic cells by macrophages is phosphatidylserine dependent
and does not induce inflammatory cytokine production. Mol. Biol. Cell 15,
1089 –1100.
56. Mathison, J. C., Tobias, P. S., Wolfson, E., Ulevitch, R. J. (1992) Plasma
lipopolysaccharide (LPS)-binding protein. A key component in macro-
phage recognition of gram-negative LPS. J. Immunol. 149, 200 –206.
57. Vassallo, R., Kottom, T. J., Standing, J. E., Limper, A. H. (2001) Vitro-
nectin and fibronectin function as glucan binding proteins augmenting
macrophage responses to Pneumocystis carinii. Am. J. Respir. Cell Mol.
Biol. 25, 203–211.
58. Gao, B., Tsan, M. F. (2003) Recombinant human heat shock protein 60
does not induce the release of tumor necrosis factor alpha from murine
macrophages. J. Biol. Chem. 278, 22523–22529.
59. Bausinger, H., Lipsker, D., Ziylan, U., Manie, S., Briand, J. P., Cazenave,
J. P., Muller, S., Haeuw, J. F., Ravanat, C., de la Salle, H., et al. (2002)
Endotoxin-free heat-shock protein 70 fails to induce APC activation. Eur.
J. Immunol. 32, 3708 –3713.
60. Gao, B., Tsan, M. F. (2003) Endotoxin contamination in recombinant
human heat shock protein 70 (Hsp70) preparation is responsible for the
induction of tumor necrosis factor alpha release by murine macrophages.
J. Biol. Chem. 278, 174 –179.
61. Reed, R. C., Berwin, B., Baker, J. P., Nicchitta, C. V. (2003) 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.
62. Lowenstein, C. J., Padalko, E. (2004) iNOS (NOS2) at a glance. J. Cell Sci.
117, 2865–2867.
63. Yonekura, H., Yamamoto, Y., Sakurai, S., Watanabe, T., Yamamoto, H.
(2005) Roles of the receptor for advanced glycation endproducts in dia-
betes-induced vascular injury. J. Pharmacol. Sci. 97, 305–311.
Rouhiainen et al. HMGB1 and cytokine expression 57
64. Sumi, D., Ignarro, L. J. (2004) Regulation of inducible nitric oxide
synthase expression in advanced glycation end product-stimulated raw
264.7 cells: the role of heme oxygenase-1 and endogenous nitric oxide.
Diabetes 53, 1841–1850.
65. Adami, C., Bianchi, R., Pula, G., Donato, R. (2004) S100B-stimulated NO
production by BV-2 microglia is independent of RAGE transducing ac-
tivity but dependent on RAGE extracellular domain. Biochim. Biophys.
Acta 1742, 169 –177.
66. Kim, J. B., Sig Choi, J., Yu, Y. M., Nam, K., Piao, C. S., Kim, S. W., Lee,
M. H., Han, P. L., Park, J. S., Lee, J. K. (2006) HMGB1, a novel
cytokine-like mediator linking acute neuronal death and delayed neuroin-
flammation in the postischemic brain. J. Neurosci. 26, 6413– 6421.
58 Journal of Leukocyte Biology Volume 81, January 2007 http://www.jleukbio.org
... The studies on cytokine-inducing properties of HMGB1 over the past years have been contradicting. We observed that rHMGB1 treated macrophages did not induce pro-inflammatory cytokines which were in consistent with the reports from other laboratories too (Rouhiainen et al., 2007;Sha et al., 2008;Tian et al., 2007;Ivanov et al., 2007;Cassetta et al., 2009). In 2007, two groups (Tian et al., 2007;Ivanov et al., 2007) reported that the activation of TLR9 by ssDNA is increased greatly in the presence of HMGB1 protein. ...
Research
Full-text available
Extracellular high mobility group box 1 (HMGB1) protein and nitric oxide (NO) has been credited with multiple inflammatory functions using in vivo and in vitro systems. Therefore, delineating their regulation may be an important therapeutic strategy for the treatment of sepsis. In the present study, it is demonstrated that recombinant HMGB1 (rHMGB1) synergizes with sub threshold concentration of TLR2 agonist (PGN; 1 g/ml) as well as with TLR4 agonist (LPS; 1 ng/ml) to induce NO release in mouse peritoneal macrophages. The enhanced iNOS expression was also observed at the transcription and translational level. Co-incubation of macrophages with rHMGB1 with either PGN or LPS showed enhanced expression of TLR2, TLR4 and RAGE. TLR2, TLR4 or RAGE knockdown macrophages effectively inhibited the rHMGB1 + PGN or LPS induced NO synergy. It was further observed that the JNK MAPK inhibitor SP600125 attenuated the PGN + rHMGB1 induced iNOS/NO synergy whereas p38 MAPK inhibitor SB908912 inhibited iNOS/NO synergy induced by LPS + rHMGB1. It was also observed that the activation of NF-B is essential for the synergy as the pharmacological inhibition or siRNA knockdown of NF-B (cRel) significantly reduced the rHMGB1 + PGN or rHMGB1 + LPS induced enhanced iNOS/NO expression. Altogether, the data suggests that the co-incubation of macrophages with rHMGB1 with either LPS or PGN induces the synergistic effect on iNOS expression and NO release by the upregulation of surface receptors (TLR2, TLR4 and RAGE) which in turn amplifies the MAPKs (p38 and JNK) and NF-B activation and results in enhanced iNOS expression and NO production.
... Heparin has a high affinity for HMGB1, which can compete with the formation of HMGB1-LPS complexes and prevent pulmonary endothelial cell pyroptosis upstream of caspase-11 activation. rHMGB1 could only induce limited cytokine secretions (Rouhiainen et al., 2007;Sha et al., 2008;Youn et al., 2008;Hreggvidsdottir et al., 2009). HMGB1 binds with other mediators, such as IL-1B, DNA, LPS, or nucleosomes, to form complexes associated with inflammation (Sha et al., 2008). ...
Article
Full-text available
Sepsis is a significant cause of mortality in critically ill patients. Acute lung injury (ALI) is a leading cause of death in these patients. Endothelial cells exposed to the bacterial endotoxin lipopolysaccharide (LPS) can progress into pyroptosis, a programmed lysis of cell death triggered by inflammatory caspases. It is characterized by lytic cell death induced by the binding of intracellular LPS to caspases 4/5 in human cells and caspase-11 in mouse cells. In mice,caspase-11-dependent pyroptosis plays an important role in endotoxemia. HMGB1 released into the plasma binds to LPS and is internalized into lysosomes in endothelial cells via the advanced glycation end product receptor. In the acidic lysosomal environment, HMGB1 permeates the phospholipid bilayer, which is followed by the leakage of LPS into the cytoplasm and the activation of caspase-11. Heparin is an anticoagulant widely applied in the treatment of thrombotic disease. Previous studies have found that heparin could block caspase-11-dependent inflammatory reactions, decrease sepsis-related mortality, and reduce ALI, independent of its anticoagulant activity. Heparin or modified heparin with no anticoagulant property could inhibit the alarmin HMGB1-LPS interactions, minimize LPS entry into the cytoplasm, and thus blocking caspase-11 activation. Heparin has been studied in septic ALI, but the regulatory mechanism of pulmonary endothelial cell pyroptosis is still unclear. In this paper, we discuss the potential novel role of heparin in the treatment of septic ALI from the unique mechanism of pulmonary endothelial cell pyroptosis.
... However, our results with HMGB1 show no significant changes in TLR4 and downstream molecules. The activation of TLR4 via HMGB1 might better work in a higher concentration or may be dependent on the complexes it forms with other molecules, immunostimulatory complexes, or with the binding of cytokines and other molecules [40,[45][46][47]. HMGB1 binds to TLR4 with remarkably less affinity than LPS, and it activates gene expression of distinct signaling patterns after stimulation. ...
Article
Full-text available
Mechanical compression simulating orthodontic tooth movement in in vitro models induces pro-inflammatory cytokine expression in periodontal ligament (PDL) cells. Our previous work shows that TLR4 is involved in this process. Here, primary PDL cells are isolated and characterized to better understand the cell signaling downstream of key molecules involved in the process of sterile inflammation via TLR4. The TLR4 monoclonal blocking antibody significantly reverses the upregulation of phospho-AKT, caused by compressive force, to levels comparable to controls by inhibition of TLR4. Phospho-ERK and phospho-p38 are also modulated in the short term via TLR4. Additionally, moderate compressive forces of 2 g/cm2 , a gold standard for static compressive mechanical stimulation, are not able to induce translocation of Nf-kB and phospho-ERK into the nucleus. Accordingly, we demonstrated for the first time that TLR4 is also one of the triggers for signal transduction under compressive force. The TLR4, one of the pattern recognition receptors, is involved through its specific molecular structures on damaged cells during mechanical stress. Our findings provide the basis for further research on TLR4 in the modulation of sterile inflammation during orthodontic therapy and periodontal remodeling.
... For recHMGB1 supply in the cell culture, 20 µg/mL recHMGB1 were coated to the plate for 1 h at room temperature and the wells were then washed by culture medium. The suspended cells were added into the wells with equal volume (4-5 × 10 5 cells) containing 10 µg/mL recHMGB1 [37,38]. For maintaining the concentration of recHMGB1 in the cell culture, the culture medium was replaced every 48 h by freshly prepared 10 µg/mL recHMGB1 in the cell culture medium. ...
Article
Full-text available
The High Mobility Group Box 1 (HMGB1) is the most abundant nuclear nonhistone protein that is involved in transcription regulation. In addition, HMGB1 has previously been found as an extracellularly acting protein enhancing neurite outgrowth in cultured neurons. Although HMGB1 is widely expressed in the developing central nervous system of vertebrates and invertebrates, its function in the developing mouse brain is poorly understood. Here, we have analyzed developmental defects of the HMGB1 null mouse forebrain, and further examined our findings in ex vivo brain cell cultures. We find that HMGB1 is required for the proliferation and differentiation of neuronal stem cells/progenitor cells. Enhanced apoptosis is also found in the neuronal cells lacking HMGB1. Moreover, HMGB1 depletion disrupts Wnt/β-catenin signaling and the expression of transcription factors in the developing cortex, including Foxg1, Tbr2, Emx2, and Lhx6. Finally, HMGB1 null mice display aberrant expression of CXCL12/CXCR4 and reduced RAGE signaling. In conclusion, HMGB1 plays a critical role in mammalian neurogenesis and brain development.
... D'autres chercheurs ont également décrit que HMGB1 hautement purifié gardait sa fonction cytokine mais des capacités pro-inflammatoires plus faibles que le HMGB1 « standard ». Ceci a également été montré pour HMGB1 extraits de cellules eucaryotes [234][235][236]. ...
Thesis
L’inflammation est le mécanisme de base du système immunitaire. Dans le cas de pathologie inflammatoire cette inflammation persiste et devient délétère pour l’organisme. Les causes de cette persistance peuvent être variées. L’une de ces causes est la présence de molécules induisant l’inflammation. Elles peuvent être d’origine exogène comme les PAMP (Pathogen-Associated Molecular Pattern). Ce sont des molécules issues des pathogènes (LPS, peptidoglycanes, ADN CpG) capables d’activer le système immunitaire. Ces molécules peuvent également être d’origine endogène comme les DAMP (Damage Associated Molecular Pattern). Ce sont des molécules libérées par les cellules en état de danger (HMGB1, HSP60) pour prévenir et activer le système immunitaire. La présence de récepteurs (TLR2, TLR4, RAGE) capable de reconnaitre ces PAMP et DAMP est également nécessaire pour pouvoir induire une inflammation. Mes travaux explorent les mécanismes moléculaires et cellulaires des PAMP et des DAMP, dans l’installation et le maintien de l’inflammation dans le cadre des maladies inflammatoires. Pour cela mon étude se focalise sur les mécanismes de reconnaissance et d’induction de l’inflammation par les PAMP et DAMP. Nous avons ainsi mis en évidence certains mécanismes cellulaires et moléculaires dans la réponse inflammatoire liés aux DAMP et PAMP. Nous nous sommes également intéressé aux récepteurs impliqués dans ces différents mécanismes et avons même mis en évidence un potentiel nouveau récepteur CD93. Nous émettons l’hypothèse que CD93 pourrait avoir un rôle dans les pathologies inflammatoires par cette capacité à lier les DAMP et PAMP.
Article
Fifty years since the initial discovery of HMGB1 in 1973 as a structural protein of chromatin, HMGB1 is now known to regulate diverse biological processes depending on its subcellular or extracellular localization. These functions include promoting DNA damage repair in the nucleus, sensing nucleic acids and inducing innate immune responses and autophagy in the cytosol and binding protein partners in the extracellular environment and stimulating immunoreceptors. In addition, HMGB1 is a broad sensor of cellular stress that balances cell death and survival responses essential for cellular homeostasis and tissue maintenance. HMGB1 is also an important mediator secreted by immune cells that is involved in a range of pathological conditions, including infectious diseases, ischaemia-reperfusion injury, autoimmunity, cardiovascular and neurodegenerative diseases, metabolic disorders and cancer. In this Review, we discuss the signalling mechanisms, cellular functions and clinical relevance of HMGB1 and describe strategies to modify its release and biological activities in the setting of various diseases.
Article
Significance: Reactive oxygen species (ROS) are well known to promote innate immune responses during and in the absence of microbial infections. However, excessive or prolonged exposure to ROS provokes innate immune signaling dysfunction and contributes to the pathogenesis of many autoimmune diseases. The relatively high basal expression of pattern recognition receptors (PRRs) in innate immune cells renders them prone to activation in response to minor intrinsic or extrinsic ROS misbalances in the absence of pathogens. Critical Issues: A prominent source of ROS are mitochondria, which are also major inter-organelle hubs for innate immunity activation, since most PRRs and downstream receptor molecules are directly located either at mitochondria or at mitochondria-associated membranes. Due to their ancestral bacterial origin, mitochondria can also act as quasi-intrinsic self-microbes that mimic a pathogen invasion and become a source of danger-associated molecular patterns (DAMPs) that triggers innate immunity from within. Recent Advances: The release of mitochondrial DAMPs correlates with mitochondrial metabolism changes and increased generation of ROS, which can lead to the oxidative modification of DAMPs. Recent studies suggest that ROS-modified mitochondrial DAMPs possess increased, persistent immunogenicity. Future Directions: Herein, we discuss how mitochondrial DAMP release and oxidation activates PRRs, changes cellular metabolism, and causes innate immune response dysfunction by promoting systemic inflammation, thereby contributing to the onset or progression of autoimmune diseases. The future goal is to understand what the tipping point for DAMPs is to become oxidized, and whether this is a road without return. Antioxid. Redox Signal. 36, 441-461.
Thesis
L'arrêt cardiaque extrahospitalier constitue un enjeu majeur de santé publique dans les pays industrialisés. Près de 40 000 cas sont par exemple dénombrés chaque année en France. Malgré l’amélioration des soins pré-hospitaliers et la reprise fréquente d’une activité cardiaque spontanée, la mortalité de ces patients reste considérable. Après la réanimation des patients, le syndrome d’ischémie-reperfusion est en effet à l’origine d’une défaillance multiviscérale associée à des dommages cérébraux sévères. Ces manifestations sont regroupées sous le terme de syndrome post-arrêt cardiaque. Ainsi, moins de 5% des patients survivent plus de 3 mois avec une bonne récupération neurologique après un arrêt cardiaque. Le principal traitement ayant fait la preuve de son efficacité clinique pour limiter le syndrome post-arrêt cardiaque consiste à assurer un contrôle ciblé de la température corporelle entre 33 et 36°C. Néanmoins, des données expérimentales et cliniques indiquent que le délai d’induction du refroidissement après la réanimation conditionne grandement son effet neuroprotecteur. Dans ce contexte, le laboratoire d'accueil étudie une stratégie susceptible d'améliorer la survie et la récupération neurologique au décours de la réanimation grâce à l’induction d’une hypothermie généralisée ultra-rapide (32-34°C). Ils proposent pour cela d’utiliser les poumons comme bio-échangeurs thermiques tout en maintenant des échanges gazeux normaux par une ventilation liquide totale (VLT) avec des perfluorocarbones. Expérimentalement, cette stratégie a démontré de puissants effets cardio-, neuro- et néphroprotecteurs dans divers modèles animaux d’ischémie généralisée. L’objectif de mes travaux de thèse était d’étudier les mécanismes sous-tendant ces effets et conditionnant la fenêtre d’efficacité de l’hypothermie après un arrêt cardiaque. Nous nous sommes plus précisément intéressés aux effets de la VLT sur la composante inflammatoire du syndrome post-arrêt cardiaque, en étudiant un élément initiateur de cette réponse, la « High Mobility Group Box 1 » (HMGB1). Nous avons tout d’abord mené une étude visant à définir les paramètres de VLT offrant le meilleur rapport tolérance/efficacité. Le laboratoire a ainsi mis au point une nouvelle méthode de « lung conservative liquid ventilation » basée sur un remplissage incomplet des poumons avec les perfluorocarbones. Nous avons ensuite évalué l’effet de la VLT sur la mortalité cellulaire précoce et l’activation immunitaire au cours de la phase aiguë du syndrome post-arrêt cardiaque. Nos travaux ont permis de montrer que, malgré l’absence de protection vis-à-vis de la libération de HMGB1 induite par l’arrêt cardiaque, la VLT permettait de moduler de façon transitoire la réponse humorale précoce qui en découle et qui contribuerait à la propagation des lésions tissulaires. Enfin, nous avons évalué le potentiel neuroprotecteur de l’inhibition pharmacologique de HMGB1 après la réanimation par une administration de glycyrrhizine. Ce traitement a également permis une réduction des séquelles neurologiques et de la défaillance hémodynamique induites par l’arrêt cardiaque. Une partie de ces effets protecteurs semblait médiée par une inhibition spécifique de la réponse lymphocytaire. Le syndrome inflammatoire consécutif à l’arrêt cardiaque apparait donc complexe et semble mettre en jeu des voies relativement spécifiques contribuant in fine la détérioration du pronostic neurologique.
Article
As clinically demonstrated by the success of immunotherapies to improve survival outcomes, tumors are known to gain a survival advantage by circumventing immune surveillance. A defining feature of this is the creation and maintenance of a tumor immune microenvironment (TIME) that directly and indirectly alters the host’s immunologic signaling pathways through a variety of mechanisms. Tumor-intrinsic mechanisms that instruct the formation and maintenance of the TIME have been an area of intensive study, such as the identification and characterization of soluble factors actively and passively released by tumor cells that modulate immune cell function. In particular, damage-associated molecular pattern molecules (DAMPs) typically released by necrotic tumor cells are recognized by innate immune receptors such as Toll-like receptors (TLRs) and stimulate immune cells within TIME. Given their broad and potent effects on the immune system, a better understanding for how DAMP and TLR interactions sculpt the TIME to favor tumor growth would identify new strategies and approaches for cancer immunotherapy.
Article
Full-text available
Endotoxin, a constituent of Gram-negative bacteria, stimulates macrophages to release large quantities of tumor necrosis factor (TNF) and interleukin-1 (IL-1), which can precipitate tissue injury and lethal shock (endotoxemia). Antagonists of TNF and IL-1 have shown limited efficacy in clinical trials, possibly because these cytokines are early mediators in pathogenesis. Here a potential late mediator of lethality is identified and characterized in a mouse model. High mobility group–1 (HMG-1) protein was found to be released by cultured macrophages more than 8 hours after stimulation with endotoxin, TNF, or IL-1. Mice showed increased serum levels of HMG-1 from 8 to 32 hours after endotoxin exposure. Delayed administration of antibodies to HMG-1 attenuated endotoxin lethality in mice, and administration of HMG-1 itself was lethal. Septic patients who succumbed to infection had increased serum HMG-1 levels, suggesting that this protein warrants investigation as a therapeutic target.
Article
Full-text available
We have recently identified two novel cysteine proteinase inhibitors from the skin of Atlantic salmon ( Salmo salar L.), named salmon kininogen and salarin. In preliminary experiments, the proteins were found to be both N- as well as O-glycosylated. In the present study we show that both proteins carry biantennary α2,3-sialylated N-glycans. A very high amount of O-acetylated Neu5Ac units are present in the N-glycans, comprising about 60% di-O-acetylated species. Non-O-acetylated Neu5Ac make up less than 5% of the sialic acids in the N-glycans. A small number of Neu5Acα2-8Neu5Ac structures were observed in the N-glycans as well. O-glycans from both proteins were recovered by reductive beta-elimination and were identified by mass spectrometric methods as mono- and disialylated core type 1 tri- and tetrasaccharides. The method used for O-glycan isolation prevented the identification of possible O-acetylation in the O-glycan-bound sialic acids, but O-acetylation was observed in one O-glycosylated peptide isolated from trypsin digest of salarin. The chemical nature of the sialic acid modifications was further studied by liquid chromatography tandem mass spectrometry of 1,2-diamino-4,5-methylenedioxybenzene–derivatized sialic acids, revealing 7-, 8-, and 9- but no 4-O-acetylation. To our knowledge, these are the first observations of sialic acid O-acetylation in N-glycans on fish species and represent clearly the most extensive N-glycan O-acetylation described on any species.
Article
A membrane-bound adhesive protein that promotes neurite outgrowth in brain neurons has been isolated from rat brain (Rauvala, H., and R. Pihlaskari. 1987. J. Biol. Chem. 262:16625-16635). The protein is an immunochemically distinct molecule with a subunit size of approximately 30 kD (p30). p30 is an abundant protein in perinatal rat brain, but its content decreases rapidly after birth. In the present study the amino-terminal sequence of p30 was determined by automated Edman degradations. A single amino-terminal sequence was found, which is not present in previously studied adhesive molecules. This unique sequence has a cluster of five positive charges within the first 11 amino acid residues: Gly-Lys-Gly-Asp-Pro-Lys-Lys-Pro-Arg-Gly-Lys. Antisynthetic peptide antibodies that recognize this sequence were produced in a rabbit, purified with a peptide affinity column, and shown to bind specifically to p30. The antipeptide antibodies were used, together with anti-p30 antibodies, to study the localization of p30 in brain cells and in neuroblastoma cells as follows. (a) Immunofluorescence and immunoelectron microscopy indicated that p30 is a component of neurons in mixed cultures of brain cells. The neurons and the neuroblastoma cells expressed p30 at their surface in the cell bodies and the neurites. In the neurites p30 was found especially in the adhesive distal tips of the processes. In addition the protein was detected in ribosomal particles and in intracellular membranes in a proportion of cells. (b) The antibodies immobilized on microtiter wells enhanced adhesion and neurite growth indicating that p30 is surface exposed in adhering neural cells. (c) Immunoblotting showed that p30 is extracted from suspended cells by heparin suggesting that a heparin-like structure is required for the binding of p30 to the neuronal cell surface. A model summarizing the suggested interactions of p30 in cell adhesion and neurite growth is presented.
Article
High-mobility-group (HMG) proteins are a family of non-histone chromosomal proteins which bind to DNA. They have been implicated in multiple aspects of gene regulation and cellular differentiation. Sulfoglucuronyl carbohydrate binding protein, SBP-1, which is also localized in the neuronal nuclei, was shown to be required for neurite outgrowth and neuronal migration during development of the nervous system. In order to establish relationship between SBP-1 and HMG family proteins, two HMG proteins were isolated and purified from developing rat cerebellum by heparin–sepharose and sulfatide-octyl–sepharose affinity column chromatography and their biochemical and biological properties were compared with those of SBP-1. Characterization by high performance liquid chromatography–mass spectrometry (HPLC–MS), partial peptide sequencing and western blot analysis showed the isolated HMG proteins to be HMG-1 and HMG-2. Isoelectric focusing, HPLC–MS and peptide sequencing data also suggested that HMG-1 and SBP-1 were identical. Similar to SBP-1, both HMG proteins bound specifically to sulfated glycolipids, sulfoglucuronylglycolipids (SGGLs), sulfatide and seminolipid in HPTLC-immuno-overlay and solid-phase binding assays. The HMG proteins promoted neurite outgrowth in dissociated cerebellar cells, which was inhibited by SGGLs, anti-Leu7 hybridoma (HNK-1) and anti-SBP-1 peptide antibodies, similar to SBP-1. The proteins also promoted neurite outgrowth in explant cultures of cerebellum. The results showed that the cerebellar HMG-1 and -2 proteins have similar biochemical and biological properties and HMG-1 is most likely identical to SBP-1.
Article
We describe methods for the isolation, purification, and characterization of full-length high-mobility group box 1 (HMGB1) and truncated mutants expressed in bacteria and in mammalian Chinese Hamster Ovary (CHO) cells. HMGB1 is an abundant nuclear and cytoplasmic protein, highly conserved across species and widely distributed in eukaryotic cells from yeast to man. As a ubiquitous nuclear DNA binding protein, HMGB1 binds DNA, facilitates gene transcription, and stabilizes nucleosome structure. In addition to these intracellular roles, HMGB1 can be released into the extracellular milieu by activated innate immune cells (i.e., macrophages, monocytes) and functions as a mediator of lethal endotoxemia and sepsis. The proinflammatory cytokine activity of HMGB1 has become an intense area of research and recombinant protein can be a useful tool to probe HMGB1 functions. Due to its dipolar charged properties, HMGB1 isolated by some methods can be contaminated with bacterial products (such as CpG DNA or lipopolysaccharide [LPS]) that may interfere with immunological analyses. Here we report our newly developed methods for the isolation and purification of biologically active HMGB1 from bacteria or mammalian CHO cells that is essentially free of contaminants. This strategy provides an important advance in methodology to facilitate future HMGB1 studies.
Article
The nonhistone chromosomal proteins, HMG-1 and HMG-2, have a folded conformation, with a high alpha-helical content, over a wide pH range. At high and low pH values, the molecules unfold. Both molecules contain cysteine and tryptophan. The tryptophans appear to be buried in the folded form. HMG-1 shows aggregation at pH 5.7, as does HMG-2 at pH 9.0. The folded form is insensitive to high concentrations of salt, suggesting that charge-charge interaction plays no role in stabilizing the tertiary structure.