Skip to main content
Erschienen in: BMC Immunology 1/2013

Open Access 01.12.2013 | Research article

Prothymosin α and a prothymosin α-derived peptide enhance TH1-type immune responses against defined HER-2/neu epitopes

verfasst von: Kyriaki Ioannou, Evelyna Derhovanessian, Eleni Tsakiri, Pinelopi Samara, Hubert Kalbacher, Wolfgang Voelter, Ioannis P Trougakos, Graham Pawelec, Ourania E Tsitsilonis

Erschienen in: BMC Immunology | Ausgabe 1/2013

Abstract

Background

Active cancer immunotherapies are beginning to yield clinical benefit, especially those using peptide-pulsed dendritic cells (DCs). Different adjuvants, including Toll-like receptor (TLR) agonists, commonly co-administered to cancer patients as part of a DC-based vaccine, are being widely tested in the clinical setting. However, endogenous DCs in tumor-bearing individuals are often dysfunctional, suggesting that ex vivo educated DCs might be superior inducers of anti-tumor immune responses. We have previously shown that prothymosin alpha (proTα) and its immunoreactive decapeptide proTα(100–109) induce the maturation of human DCs in vitro. The aim of this study was to investigate whether proTα- or proTα(100–109)-matured DCs are functionally competent and to provide preliminary evidence for the mode of action of these agents.

Results

Monocyte-derived DCs matured in vitro with proTα or proTα(100–109) express co-stimulatory molecules and secrete pro-inflammatory cytokines. ProTα- and proTα(100–109)-matured DCs pulsed with HER-2/neu peptides induce TH1-type immune responses, prime autologous naïve CD8-positive (+) T cells to lyse targets expressing the HER-2/neu epitopes and to express a polyfunctional profile, and stimulate CD4+ T cell proliferation in an HER-2/neu peptide-dependent manner. DC maturation induced by proTα and proTα(100–109) is likely mediated via TLR-4, as shown by assessing TLR-4 surface expression and the levels of the intracellular adaptor molecules TIRAP, MyD88 and TRIF.

Conclusions

Our results suggest that proTα and proTα(100–109) induce both the maturation and the T cell stimulatory capacity of DCs. Although further studies are needed, evidence for a possible proTα and proTα(100–109) interaction with TLR-4 is provided. The initial hypothesis that proTα and the proTα-derived immunoactive decapeptide act as “alarmins”, provides a rationale for their eventual use as adjuvants in DC-based anti-cancer immunotherapy.
Begleitmaterial
Hinweise

Electronic supplementary material

The online version of this article (doi:10.​1186/​1471-2172-14-43) contains supplementary material, which is available to authorized users.

Competing interests

The authors declare that they have no competing interests.

Authors’ contributions

KI: performed the experiments, analyzed data, carried out statistical analyses and wrote the manuscript. ED: designed, analysed and interpreted flow cytometry data and helped to write the manuscript. ET: participated in immunoblotting data acquisition and analyses. PS: performed sample collection and helped to draft the manuscript. HK: carried out peptide synthesis and purification and helped to draft the manuscript. WV: helped in HLA-typing and to draft the manuscript. IPT: participated in the design of the study and reviewed the manuscript. GP: participated in the design and coordination of the study, helped to draft, reviewed and edited the manuscript. OET: conceived, designed and coordinated the study, drafted and reviewed the manuscript. All authors read and approved the final manuscript.
Abkürzungen
HMGB1
High mobility group box 1
HSP
Heat shock protein
MFI
Mean fluorescence intensity
proTα
Prothymosin alpha
TLR
Toll-like receptor.

Background

Anti-cancer vaccines are designed to break tolerance to self and stimulate strong and durable anti-tumor immunity. Administering defined tumor-derived epitopes to cancer patients for the activation of helper and cytotoxic T cells has been shown to enhance anti-cancer immune responses in vivo and in some cases to lead to objective clinical responses [13]. To optimize the efficacy of peptide-based anti-cancer vaccines, combinatorial approaches stimulating both innate and adaptive immunity are now being clinically evaluated [4, 5]. Mature dendritic cells (DCs) are key players for eliciting such responses, as they present antigens to T cells and provide the necessary co-stimulatory signals and cytokines favoring the efficient activation of tumor-reactive immune cells [6, 7]. DC maturation can be induced in vivo upon admixing and co-administering immunogenic peptides with adjuvants, but to date this strategy has been proven successful only when vaccinating against common pathogens [8]. In cancer patients, the presence of tumor-associated suppressive factors impairs endogenous DC functions [9], a condition that can be bypassed only by the adoptive transfer of ex vivo matured immunocompetent DCs [10, 11].
Adjuvants comprise, among others, Toll-like receptor (TLR) agonists, the majority of which reportedly promotes DC maturation [12]. A subcategory thereof are molecules with so-called pathogen-associated molecular patterns (PAMPs), such as CpG oligodeoxynucleotides that signal through TLR-9 [13], poly-I:C ligating TLR-3 [14], imiquimod, a TLR-7 agonist [15] and monophosphoryl lipid A, a TLR-4 agonist [16]. A second group consists of molecules possessing damage-associated molecular patterns (DAMPs) or “alarmins”. High mobility group box 1 (HMGB1) protein and heat shock protein (HSP) 90 are notable examples of DAMPs. Both proteins are strictly intracellular under normal physiological conditions, but when excreted eg. from damaged cells, signal through TLR-4, sensitize DCs and promote adaptive immune responses [17]. This functional dualism, in and out of the cell, also characterizes prothymosin alpha (proTα).
In normal living cells, proTα is localized in the nucleus where it controls the cell cycle and promotes cell proliferation. Released from dead cells, extracellular proTα acquires multi-functional immunomodulatory properties [18]. We and others have previously shown that proTα upregulates the expression of IRAK-4 in human monocytes [19], ligates TLR-4 on murine macrophages and signals through MyD88-dependent and independent pathways [20]. Similar to its immunoreactive decapeptide proTα(100–109) [21], it upregulates the expression of HLA-DR [22], CD80, CD83 and CD86 and promotes maturation of human DCs in vitro[23].
Here, we show that DCs matured ex vivo in the presence of proTα or proTα(100–109) are not only phenotypically but also functionally competent, secrete pro-inflammatory cytokines and induce TH1-type immune responses in the presence of tumor-associated immunogenic epitopes of the oncoprotein HER-2/neu. DCs matured with proTα or proTα(100–109) prime naïve CD8-positive (+) T cells to exert HER-2/neu peptide-specific cytotoxicity and CD4+ T cells to proliferate in a peptide-dependent manner. Finally, we provide preliminary evidence suggesting that both proTα and its decapeptide proTα(100–109) likely signal via TLR-4 in human DCs.

Results

Phenotype of and cytokine production by proTα- or proTα(100–109)-matured DCs

We have previously shown that proTα and proTα(100–109) efficiently mature human DCs in vitro, as indicated by the induction of surface expression of established DC-markers to levels comparable to those induced by lipopolysacharide (LPS) [23] or tumor necrosis factor (TNF)-α (this report). As shown in Figure 1, LPS-matured DCs significantly upregulated the expression of HLA-DR, CD11b, CD80, CD83, CD86 and CD40, to levels comparable to TNF-α-matured DCs (p < 0.001 compared to iDCs for all values). Similarly, both agents caused a reduction of CD14 expression. In agreement with our previous study [23], iDCs matured with either proTα or its decapeptide presented a similar phenotype to LPS- or TNF-α-matured DCs, upregulating the expression of all co-stimulatory molecules (p < 0.05-0.001, compared to iDCs; Figure 1) and downregulating CD14 (p < 0.01, compared to iDCs).
In conjunction with their phenotype, functionally competent mature DCs secrete pro- and anti-inflammatory cytokines [6]. Therefore, we assessed the production of TNF-α, interleukin (IL)-12 and IL-10 from iDCs and mature DCs and determined the IL-12:IL-10 ratio. We present these data because high TNF-α levels, as well as a balance between IL-12:IL-10 in favor of IL-12 have been proposed to promote TH1-polarization [24]. As shown in Figure 2, iDCs produced low amounts of all three cytokines, whereas mature DC supernatants collected 48 h after addition of LPS or TNF-α were rich in TNF-α and IL-12. Compared to iDCs, higher TNF-α and IL-12 levels were also found in supernatants of cultures of DCs matured with proTα and proTα(100–109). Although the absolute concentrations of cytokines varied among the differentially matured DCs, their overall cytokine-production patterns were comparable. Most importantly, the mean IL-12:IL-10 ratios were similar (6.61, 6.45, 7.89 and 5.18, for LPS-, TNF-α-, proTα- and proTα(100–109)-matured DCs, respectively). These data suggest that the peptides bias immunoreactivity towards a pro-inflammatory TH1-type of response.
Finally, in the presence of a blocking antibody against TLR-4 (a-TLR-4; Figure 2), lower amounts of cytokines were secreted by LPS-, proTα- and proTα(100–109)-matured DCs, but not TNF-α-matured DCs. Notably, a-TLR-4 reduced the levels of LPS- and proTα-induced IL-12 production by 55 and 47%, respectively (p < 0.05), implying that IL-12 production by LPS- and proTα-matured DCs is at least partially, TLR-4-dependent [25].

ProTα and proTα(100–109) lead to TH1-polarized tumor peptide-reactive immune response

As optimally matured DCs prime antigen-specific CD4+ and CD8+ T cell activation and proliferation of naive T cells [7], we next assessed whether proTα- and proTα(100–109)-matured DCs are functionally competent, i.e., induce in vitro the selective expansion of tumor antigen-specific T cells.
Monocyte-derived DCs matured for 48 h with proTα, proTα(100–109) or TNF-α (used as a conventional DC maturation agent) were pulsed with the HLA-A2 and HLA-DR4-restricted HER-2/neu(369–377) [HER-2(9369)] and HER-2/neu(776–790) [HER-2(15776)] epitopes, and used to prime autologous naïve T cells isolated from the peripheral blood of HLA-A2+/DR4+ healthy donors. T cells were restimulated twice, at weekly intervals, with similarly matured autologous DCs. Twelve hours after the third stimulation their production of TNF-α, interferon (IFN)-γ, IL-2, IL-4, IL-10 and IL-17 was analysed. Figure 3 shows the percentages of IFN-γ+, IL-2+, IL-4+ and IL-10+ CD4+ T cells from one representative donor of 5 tested with similar results (Additional file 1: Table S1A). In the presence of unpulsed TNF-α-matured DCs, only a low percentage of CD4+ T cells produced IFN-γ (0.02%), which was significantly increased (23.30%) in the presence of the HER-2/neu peptides. An analogous increase in the percentage of IFN-γ-producing cells was also recorded in CD4+ T cells stimulated with proTα- or proTα(100–109)-matured DCs in the presence of the same peptides (21.78% and 22.93%, respectively, compared to 0.01% and 0.02% in the absence of HER-2/neu peptides; Figure 3). The percentages of IL-2-producing CD4+ T cells in all groups were also significantly enhanced upon stimulation with HER-2/neu peptide-pulsed DCs (35.02% for TNF-α-, 31.51% for proTα- and 37.18% for proTα(100–109)-matured DCs, compared to 0.08%, 0.09% and 0.71% of the respective unpulsed groups; Figure 3). A similar enhancement was also seen for TNF-α-producing CD4+ T cells (Additional file 1: Table S1A). In contrast, production of IL-4 and IL-10 was only marginally increased when CD4+ T cells were stimulated with HER-2/neu peptide-pulsed DCs, regardless of the factor used for DC maturation (Figure 3). Specifically, the percentage of CD4+ T cells producing IL-4 was increased from 0.01% (without peptides) to 1.17% among T cells stimulated with peptide-pulsed TNF-α-matured DCs, from 0.01% to 0.61% in the proαα- and from 0.02% to 0.82% in the proTα(100–109)-matured DC groups. A minor enhancement was also recorded for IL-10 production; IL-10+ cells increased from 0.03% to 0.11% in the TNF-α-, from 0.01% to 0.02% in the proαα- and from 0.01% to 0.02% in the proTα(100–109)-matured DC cultures without and with HER-2/neu peptides, respectively. IL-17 production exhibited a similar pattern of marginal increase in CD4+ T cells stimulated in the presence of all matured antigen-pulsed DCs (Additional file 1: Table S1A). These data suggest that proTα- and proTα(100–109)-matured DCs are immunocompetent and in the presence of specific tumor antigenic epitopes, favor the in vitro production of pro-inflammatory (IFN-γ, IL-2, TNF-α), rather than anti-inflammatory cytokines (IL-4, IL-10) and IL-17 by CD4+ T cells, inducing TH1-type immune responses.

ProTα- and proTα(100–109)-matured DCs stimulate tumor peptide-specific CD8+ T cell responses

Cell-mediated immunity requires initial collaboration between TH1 CD4+ and CD8+ T cells [26]. Thus, we next investigated whether proTα- and proTα(100–109)-matured DCs can elicit tumor peptide-specific cytotoxic T cell immune responses.
CD8+ T cells recovered from the same stimulation cultures as aforementioned were assessed for the intracellular production of TNF-α. As shown in Figure 4A, they also exhibited a similar pattern of enhanced cytokine production in the presence of HER-2/neu peptides as did CD4+ T cells. The percentage of TNF-α+ cells was increased from 0.35% (unpulsed) to 47.52% (pulsed) when T cells were stimulated with TNF-α-matured DCs, and from 0.12% to 45.38% for proTα- and from 0.13% to 42.88% for proTα(100–109)-matured DCs (Figure 4A). In addition and in accordance with the results recorded for CD4+ T cells, IL-2- and IFN-γ-producing CD8+ T cells were also increased in the presence of peptide-pulsed DCs in the cultures, whereas differences in the percentages of IL-10-producing CD8+ T cells were only marginal (Additional file 1: Table S1A).
The same cells were assessed for the expression of CD107a, as a surrogate marker for cytotoxicity [27]. In the absence of HER-2(9369), a low percentage of CD8+ T cells stimulated with TNF-α-matured DCs expressed CD107a (3.70%; Figure 4A), which increased when cells were stimulated with HER-2(9369)-pulsed DCs (54.75%). Similar CD107a upregulation was observed in CD8+ T cells stimulated with proTα- and proTα(100–109)-matured HER-2(9369)-pulsed DCs (36.86% and 41.99%, respectively, compared to 2.80% and 2.17% of the unpulsed groups; Figure 4A). Since TNF-α mediates target cell damage and CD107a-expressing CD8+ T cells are cytotoxic [27], our results suggest that proTα- and proTα(100–109)-matured DCs efficiently activate CD8+ cytotoxic T cells, which were able to kill targets presenting the immunogenic epitope versus which they were primed.
Cytotoxic activity was verified by using 51Cr-labeled HLA-A2+ T2 cells loaded with HER-2(9369) or an irrelevant epitope, tyrosinase(369–377) [tyr(9369)]. CD8+ T cells thrice stimulated with peptide-pulsed TNF-α-, proTα- or proTα(100–109)-matured DCs were coincubated with these peptide-loaded T2 targets. The results showed that CD8+ T cell mean cytotoxicity against non-peptide loaded T2 targets did not exceed 30% in any group (26.9% for TNF-α, 23.7% for proTα- and 21.4% for proTα(100–109)-matured DCs; Figure 4B), whereas HER-2(9369)-loaded T2 targets were lysed twice as efficiently by CD8+ T cells recovered from all stimulation cultures (49.9% for TNF-α-, 46.6% for proTα- and 40.4% for proTα(100–109)-matured DCs; Figure 4B). Cytotoxicity against T2 targets loaded with tyr(9369) was low and in no instance exceeded 30%. These cytotoxic responses were significantly decreased by monoclonal antibody (mAb) to MHC class I molecules, suggesting that the CD8+ T cells generated by our stimulation protocol are MHC class I-restricted and HER-2(9369)-specific (Figure 4B).

Polyfunctionality of HER-2(9369)–specific CD8+ T cells

Based on previous studies associating T cell polyfunctionality with high IFN-γ production and the quality of the elicited responses [28, 29], we carried out a functional analysis of the HER-2(9369)-specific CD8+ T cells generated in these experiments. Using FlowJo software, we analyzed their ability to produce effector cytokines (IFN-γ, TNF-α and IL-2) and to degranulate (expression of CD107a). Quantifying the fraction of the responsive CD8+ T cells producing any one (1+), any two (2+), any three (3+) or all four (4+) mediators, we observed that approximately a mean ~16% of the responsive CD8+ T cells were 2+ cells, regardless of the agent used to mature the DCs that stimulated them (16.26% for TNF-α; 16.92% for proTα; 15.95% for proTα(100–109); Figure 5). In all experimental groups, 3+ cells were also detected in increased percentages (8.13% for TNF-α; 7.03% for proTα; 4.34% for proTα(100–109)). In contrast, very few 4+ cells were detected under any conditions. Taken together, these data suggest that proTα- or proTα(100–109)-matured DCs were able to induce polyfunctional (2+, 3+) CD8+ peptide-specific T cell responses at least as well as TNF-α-matured DCs.

T cells stimulated with proTα- or proTα(100–109)-matured DCs proliferate in response to the HER-2(15776) epitope

T cell proliferation was assessed by the incorporation of 3H-thymidine. HER-2(15776)-sensitized T cells coincubated with HER-2(15776)-, tyrosinase(448–462) [tyr(15448)]-pulsed or unpulsed DCs, specifically proliferated in response only to the HER-2/neu epitope (Table 1). ProTα- and proTα(100–109)-matured DCs showed relatively high mean stimulation indices (S.Is) (2.64 and 2.26, respectively), comparable to those recorded for TNF-α-matured DCs (3.09). Addition of mAb to MHC class II molecules reduced mean S.Is in all groups (0.87 for TNF-α; 0.93 for proTα; 0.98 for proTα(100–109)). These results suggest that following our in vitro culture protocol, peptide-reactive T cells are generated, which proliferate only in a HER-2(15776)-dependent MHC class II-restricted manner.
Table 1
T cells stimulated with proTα- or proTα(100–109)-matured DCs proliferate in the presence of HER-2(15 776 )-pulsed DCs
DCs matured with
DCs pulsed with
Mean counts per minute (cpm) ± SD*
Stimulation index (S.I.) ± SD*
TNF-α
-
13693 ± 1413
1
HER-2(15776)
42314 ± 7139
3.09 ± 0.52
tyr(15448)
16433 ± 1840
1.20 ± 0.13
HER-2(15776) + anti-MHC class II
11914 ± 2033
0.87 ± 0.15
ProTα
-
13145 ± 1742
1
HER-2(15776)
34702 ± 5143
2.64 ± 0.39
tyr(15448)
17220 ± 2974
1.31 ± 0.23
HER-2(15776) + anti-MHC class II
12225 ± 2603
0.93 ± 0.20
ProTα(100–109)
-
14577 ± 1041
1
HER-2(15776)
32944 ± 6567
2.26 ± 0.45
tyr(15448)
15306 ± 3608
1.05 ± 0.25
HER-2(15776) + anti-MHC class II
14285 ± 2989
0.98 ± 0.20
* Mean cpm, S.I. ± SD from 2–5 donors.

ProTα and proTα(100–109) induce the maturation of DCs via triggering TLR-4

We have previously reported that stimulation of human monocytes with proTα upregulated IRAK-4, a protein kinase involved in TLR downstream signaling [19], whereas Mosoian et al. [20] showed that proTα ligates TLR-4 and signals through both TRIF- and MyD88-dependent pathways. To determine whether TLR-4 is triggered by our peptides, we studied the kinetics of TLR-4 surface expression on proTα and proTα(100–109)-stimulated DCs. Immature DCs (iDCs) and DCs matured with LPS (a known TLR-4 ligand; [30]), proTα or proTα(100–109) for 15 min, 30 min, 1 h, 18 h and 36 h were analyzed by flow cytometry. The percentage of surface TLR-4 expression over time is presented in Figure 6. Maturation of DCs with LPS led to an early (15 and 30 min) decrease of TLR-4 expression (by ~15%) due to internalization [31], and a subsequent increase from 1 to 18 h [32]. At 36 h post-LPS addition, TLR-4 expression was lower and comparable to that of iDCs (0 h). ProTα and proTα(100–109), marginally downregulated TLR-4 expression at 30 min and, similarly to LPS, transiently increased it from 1 to 18 h. As with LPS-matured DCs, basal levels of TLR-4 expression were detected at 36 h post-maturation. The similar kinetics of TLR-4 expression in LPS-, proTα- and proTα(100–109)-matured DCs is consistent with the notion that the two peptides interact with TLR-4.
To extend these findings, we next investigated the intracellular expression levels of three adaptor molecules that participate in signaling pathways downstream of TLR-4, namely TRIF, an adaptor molecule common to TLR-3 and -4 signaling; TIRAP, a signaling adaptor common to TLR-2, and -4; and MyD88, a molecule upregulated upon ligation of all TLRs except TLR-3 [33]. We specifically selected these three adaptors because this constellation is unique to TLR-4 activation. Total cell extracts from iDCs and DCs matured with LPS, proTα or proTα(100–109) for 1 h and 18 h were immunoblotted (Figure 7A). Upon densitometric quantification of each protein band detected, expression relative to GAPDH was calculated. As shown in Figure 7B, addition of LPS led to a significant ~2-3 fold increase of the expression of all three adaptors within 1 h (3.05 for TRIF, 2.88 for TIRAP and 1.81 for MyD88) relative to iDCs (1.38 for TRIF, 1.00 for TIRAP and 0.74 for MyD88). At 18 h post-addition of LPS, expression of all adaptors was decreased and again comparable to iDCs (1.35 for TRIF, 1.16 for TIRAP and 1.20 for MyD88). A similar trend of increased expression was also observed 1 h after addition of proTα (2.537 for TRIF, 2.28 for TIRAP and 1.577 for MyD88) or proTα(100–109) (1.62 for TRIF, 1.423 for TIRAP, 1.06 for MyD88), although in the latter case, the detected protein levels were lower. As with LPS, 18 h after proTα or proTα(100–109) DC-stimulation, the expression of TRIF, TIRAP and MyD88 was reduced and was similar to iDCs. These data, in conjunction with the cytokine profile shown in Figure 2, suggest that LPS, proTα, and possibly also proTα(100–109), activate DCs at least partly through one common TLR-4-dependent intracellular signaling pathway.

Discussion

We have previously shown that human monocyte-derived iDCs activated in vitro with proTα or its immunoreactive decapeptide, proTα(100–109), acquire a mature DC phenotype [23]. Here, we show that DC maturation induced by proTα or proTα(100–109) promotes the secretion of IL-12, rather than IL-10, from these cells. Thus, both proTα- and proTα(100–109)-matured DCs possess immunostimulatory properties appropriate for the efficient activation of T cells, through their enhanced antigen-presenting capacity (HLA-DR; signal 1), the increased expression of co-stimulatory molecules (CD80/CD86; signal 2) and the secretion of inflammatory mediators (IL-12), recently proposed to act as signal 3 for optimizing effector T cell functions [34, 35].
We assessed whether these ex vivo generated DCs can present tumor-associated immunogenic peptides to autologous T cells, along with the appropriate signals for their activation. We pulsed DCs with one MHC class I- and one class II-restricted immunodominant epitope from the oncoprotein HER-2/neu, HER-2(9369) and HER-2/neu(15776), respectively [36, 37]. Our results show that proTα- or proTα(100–109)-matured HER-2/neu peptide-pulsed DCs favor the generation of TH1-type immune responses in vitro, by polarizing CD4+ T cells to produce pro-inflammatory cytokines. This cytokine milieu, characterized by high levels of IFN-γ and IL-2, results in the generation of strong CD8+ T cell responses [26, 38], as we also observed. Indeed, CD8+ effectors recovered from the same stimulation cultures exhibited a pro-inflammatory cytokine profile similar to the CD4+ T cells (Additional file 1: Tables S1A and B) and enhanced HER-2(9369)-specific MHC class I-restricted cytotoxicity. Of interest, a high percentage of the peptide-specific CD8+ T cells generated in our stimulation cultures were polyfunctional, a quality reportedly associated with superior T cell performance [28, 29, 39]. These findings, in conjunction with the observed enhancement of HER-2(15776)-specific T cell proliferation, suggest that in the presence of tumor antigenic peptides, proTα- and proTα(100–109)-matured DCs efficiently promote the expansion of peptide-specific T cells.
Different DC-stimulating agents, including TLR ligands, have long been and still are being explored to optimize the immunostimulatory properties of DCs [10, 11, 40, 41]. Although it was initially proposed that TLRs recognized only PAMPs, accumulating evidence to date suggests that TLRs also bind and respond to endogenous ligands released during tissue injury and inflammation, termed DAMPs or “alarmins” [42]. Most prominent among the alarmins are HMGB1, members of the HSP family and granulysin [43], all of which mature and activate DCs in vitro and bias immune responses towards a TH1-type, when used as vaccine adjuvants in vivo[4448]. We and others have previously shown that proTα promotes antigen-specific adaptive immune responses [20, 4952] and based on the data presented herein, we now identify proTα as an alarmin. Moreover, in line with data on immunoreactive peptide-fragments derived from either HMGB1 (Hp91; [53]) or HSP70 (HSP70359-610; [46]), we show that the immunologically active site of proTα, the decapeptide proTα(100–109) [23], also favors TH1-polarization and induces HER-2/neu peptide-specific immune responses.
To suggest a possible molecular mechanism underlying the effect of proTα and proTα(100–109), and considering recent data from ourselves and others [19, 20], we investigated whether TLR-4 expressed on human mature DCs is triggered by proTα or proTα(100–109). Our results show that proTα- or proTα(100–109)-induced DC maturation was associated with modulation of TLR-4 surface expression. Moreover, the expression of three TLR-4-associated intracellular adaptors, TRIF, TIRAP and MyD88, was promptly (at 1 h post-stimulation) increased in proTα- or proTα(100–109)-matured DCs, providing indirect evidence that the adjuvant activity of proTα and proTα(100–109) most likely involves TLR-4. Our data are in agreement with those of Mosoian et al. [20], showing that in murine macrophages proTα signals through the MyD88- and the TRIF-dependent pathways inducing TNF-α and type I IFN production, respectively. TLR ligation is a common mechanism of action, shared by different DAMPs. TLR-2 and -4 are involved in HMGB1 signaling in vitro[5456], and several HSPs, including HSP22, HSP60, HSP70 and HSP90 also act as TLR-4 agonists [17, 5759]. Our results add to these observations, suggesting that both proTα and its shorter immunoactive decapeptide likely signal through TLR-4. The ambiguities raised as to whether proTα and proTα(100–109) share a common mechanism of action on DCs with LPS, could be attributed to: (1) inadequate internalization of TLR-4 by monocyte-derived human DCs, which reportedly are CD14low (Figure 1, Additional file 2: Figure S1; [60, 61]). Indeed, stimulation of CD14high human monocytes and monocyte-derived human macrophages (Additional file 3: Figure S2) with proTα or proTα(100–109), induced the rapid CD14-dependent endocytosis of TLR-4, with kinetics similar to the response to LPS (Additional file 2: Figure S1); (2) differential requirements for TLR-4-mediated signaling depending on the cell population (eg. monocytes, macrophages versus DCs; [62]) and/or cell origin (eg. mouse versus human; [63]); and (3) the involvement of other TLRs (eg. TLR-2) and/or PRRs in proTα- and proTα(100–109)-induced DC signaling. In support of the latter, a similar phenomenon has been described for HMGB1; the intact protein signals through TLR-2 and -4 [53], and its immunostimulatory peptide Hp91 acts via TLR-3 or even other receptors [45].

Conclusion

Taken altogether, we show herein that proTα and proTα(100–109) optimize immunogenic peptide-pulsed DC functionalities in vitro, possibly by TLR-4 triggering. Ex vivo education of DCs by proTα or proTα(100–109) results in their polarization to type-1 DCs, with increased capacity to stimulate tumor peptide-specific T cell responses and to render cytotoxic T cells polyfunctional. If this holds true also in vivo, then these molecules could be promising components of DC-based anti-cancer vaccines.

Methods

Peptide synthesis

ProTα(100–109), and the tumor antigen epitopes HER-2(9369), tyr(9369) (HLA-A2-restricted) [64], HER-2(15776) and tyr(15448) (HLA-DR4-restricted) [36, 64] were synthesized by the Fmoc (9-fluorenylmethoxycarbonyl)/tBu chemistry utilizing a multiple peptide synthesizer Syro II (MultiSynTech, Witten, Germany). Crude peptides were purified by HPLC on a reverse phase C18 Nucleosil 100-5C column (HPLC Technologies, UK) to a purity of >95%, using a linear gradient of 5.8% acetonitrile in 0.05% trifluoroacetic acid for 45 min. All peptides were characterized by matrix-assisted laser desorption ionization-time of flight mass spectrometry and results were in all cases in agreement with the calculated masses. Human recombinant proTα was purchased from Alexis Biochemicals, CA, USA and passed through an Endotoxin removal column (Pierce Biotechnology). Prior to their use, all peptides and proTα were tested for endotoxin levels using the LAL chromogenic Endotoxin Quantitation kit (Pierce Biotechnology, IL, USA) according to the manufacturer’s instructions. They were endotoxin-free.

Cell lines and PBMC isolation

Human T2 cells (HLA-A*0201) were cultured in RPMI 1640, supplemented with 10% heat-inactivated fetal bovine serum (FBS), 2 mM L-glutamine, 10 mM Hepes, 5 μg/mL Gentamycin, 10 U/mL Penicillin and 10 U/mL Streptomycin (all from Lonza, Cologne, Germany), at 37°C, in a humidified 5% CO2 incubator.
Buffy coats were collected from HLA-A2+ and DR4+ healthy blood donors. Prior to blood draw, individuals gave their informed consent according to the regulations approved by the 2nd Peripheral Blood Transfusion Unit and Haemophilia Centre, ‘Laikon’ General Hospital Institutional Review Board, Athens, Greece. PBMCs were isolated by centrifugation over Ficoll-Histopaque (Lonza) density gradient, resuspended in X-VIVO 15 (Lonza) or cryopreserved in FBS-10% DMSO (Sigma-Aldrich Chemical Co., St Louis, MO, USA) for later use.

DC maturation and T cell stimulation

Highly enriched monocytes (>80% CD14+) were obtained from PBMCs by plastic adherence for 2 h at 37°C [65]. Non-adherent cells were removed and cryopreserved. Monocytes were cultured for 5 days in X-VIVO 15 supplemented with 800 IU/mL recombinant human granulocyte macrophage colony-stimulating factor (GM-CSF) and 500 IU/mL recombinant human IL-4 (both from R&D Systems GmbH, Wiesbaden-Nordenstadt, Germany). On day 5, iDCs were treated with LPS (0.5 μg/mL; Sigma-Aldrich), TNF-α (10 ng/mL; R&D Systems), proTα (160 ng/mL) or proαα(100–109) (25 ng/mL) for 1–48 h, concentrations already reported to induce DC maturation [23]. Mature DCs were recovered at various time points for phenotypic and TLR-4 analysis by flow cytometry and immunoblotting, or were used to stimulate autologous T cells. Supernatants from 48 h matured DCs were also collected and the concentrations of TNF-α, IL-10 and IL-12 were quantified using commercially available ELISA kits (all from Life Technologies Corporation, Carlsbad, USA), according to manufacturer’s instructions. For TLR-4 neutralization experiments, iDCs were pre-incubated in the presence of anti-TLR-4 (a-TLR-4) neutralizing monoclonal antibody (mAb; clone W7C11) or an irrelevant mouse IgG1 mAb (both from InvivoGen, San Diego, USA) at a final concentration of 10 μg/mL for 1 h and further stimulated with LPS, proTα or proTα(100–109) for 48 h. TNF-α, IL-10 and IL-12 were determined in culture supernatants.
For T cell stimulation, 48 h matured DCs (1×106/mL) were pulsed with 50 μg/mL HER-2(9369) and HER-2(15776) for 6 h at 37°C, in a humidified 5% CO2 incubator in X-VIVO 15. DCs were washed twice, resuspended in X-VIVO 15 and added to autologous lymphocytes (non-adherent fraction) at a DC:lymphocyte ratio of 1:10. T cells were stimulated thrice at weekly intervals and on days 3 and 5 after each stimulation, 40 IU/mL IL-2 (Proleukin; Novartis Pharmaceuticals Ltd, UK) were added to the cultures. At the third stimulation, Golgi-Plug (1 μL/mL; Becton-Dickinson (BD) Biosciences, Erembodegem, Belgium) was added in the cultures, and 12 h later, T cells were harvested and analyzed for cytokine production by flow cytometry.

Flow cytometry analysis

For DC phenotype analysis, iDCs and mature DCs were stained for the surface molecules HLA-DR, CD80, CD83, CD86, CD11b, CD40 and CD14. Triple staining was performed using appropriate combinations of FITC-, PE- or PE-Cy5-labelled mouse anti-human IgG1 and IgG2 mAbs (BD Biosciences) at saturating concentrations for 30 min on ice. DCs were also stained with irrelevant anti-human IgG1 and IgG2 mAbs (BD Biosciences), as isotype controls. Samples were measured using a FACSCalibur flow cytometer (BD Biosciences) and data were analyzed using CellQuest software. MFI was evaluated for each marker.
For TLR-4 expression, iDCs and DCs matured with LPS, proTα or proTα(100–109) for 15 min, 30 min, 1 h, 18 h and 36 h were harvested and treated with human immunoglobulin (GAMUNEX; Bayer, Leverkusen, Germany) and ethidium monoazide (EMA; Invitrogen, Karlsruhe, Germany) to block Fc receptors and label nonviable cells, respectively. DCs were then stained with TLR-4/Brilliant Violet 421, CD11c/PE-Cy7 (both from BioLegend, San Diego, CA) and Lineage 1 cocktail/FITC (BD Biosciences) mAbs and measured immediately using LSR II or FACSCanto II and FACSDiva software (BD Biosciences). Data were analyzed using FlowJo software (TreeStar, Ashland, OR). Duplicates were excluded using the forward-scatter area versus forward-scatter height plot, TLR-4+ cells were gated within viable DCs (EMA-negative (−), CD11c + and Lineage 1-) and their MFI was determined. For TLR-4 neutralization experiments, a-TLR-4-treated iDCs were stimulated as above and stained with CD14/FITC (BioLegend) and TLR-4/Brilliant Violet 421 or PE (BioLegend) mAbs at saturating concentrations for 30 min on ice. DCs were also stained with irrelevant anti-human IgG2 mAbs (BD Biosciences), as isotype controls. Samples were measured using a FACSCanto II and data were analyzed using FACSDiva software.
For cytokine production analysis, T cells were harvested and treated with GAMUNEX and EMA. They were then stained with the following mAbs: CD3/eFluor 605, IL-10/PE, and IL-17/PerCP-Cy5.5 (eBioscience, San Diego, CA); CD-4/PerCP, CD-8/APC-H7, IL-4/APC, IFN-γ/PE-Cy7 and CD107a/FITC (BD Biosciences); IL-2/Alexa700 and TNF-α/Brilliant Violet 421 (BioLegend). Samples were analysed immediately using an LSR II and FACSDiva software and data were processed using FlowJo software. Duplicates were excluded using the forward-scatter area versus forward-scatter height plot, and CD4+ and CD8+ cells were gated within viable CD3+ lymphocytes and analyzed separately for cytokine production. The percentage of cells producing each cytokine on gated T cells was determined.

Cytotoxicity assay

The cytotoxic activity of thrice stimulated T cells was determined by standard 51Cr- release assay. T2 cells were incubated for 2 h at 37°C with 10 μg/mL HER-2(9369) or tyr(9369), washed and labeled with sodium chromate, as previously described [21]. Non-loaded T2 were similarly labeled for controls. Effectors (1×106/mL in X-VIVO 15; 100 μL/well) were seeded in 96-well U-bottom plates (Greiner Bio-one, Kirchheim, Germany) and T2 targets were added (5×104/mL; 100 μL/well), at an effector:target (E:T) ratio of 10:1. Where indicated, mAb to MHC class I molecules (W6/32, kindly donated by Prof. S. Stevanovic, University of Tübingen) was added to the cultures at a final concentration of 5 μg/mL for the entire incubation period [66]. After 18 h of coincubation at 37°C, 5% CO2, 100 μL of supernatant were removed from each well and isotope (counts per minute (cpm)) was counted in a γ-counter (1275 Mini-gamma LKB Wallac, Turku, Finland). To determine maximal and spontaneous isotope release, targets were incubated with 3 N HCl and in plain medium, respectively. All cultures were set in triplicate. Percentage of specific cytotoxicity was calculated according to the formula: [(cpm experimental-cpm spontaneous)/(cpm maximal-cpm spontaneous)] ×100.

Proliferation assay

Stimulated T cells were seeded in 96-well U-bottom plates (1 × 106/mL; 100 μL). Autologous matured DCs pulsed with 50 μg/mL HER-2(15776) or tyr(15448) for 6 h, were added (1 × 105/mL; 100 μL/well) and cocultured for 5 days. T cells incubated with unpulsed matured DCs or in the presence of IL-2 (500 IU/mL) were used as controls. Where indicated, mAb to MHC class II molecules (L243, kindly donated by Prof. S. Stevanovic) was added to the cultures at a concentration of 5 μg/mL for the entire culture period [66]. For the last 18 h of culture, 1 μCi 3H-thymidine (Amersham Pharmacia Biotech, Amersham, Bucks, UK) was added per well and cells were harvested in a semi-automatic cell harvester (Skatron Inc., Tranby, Norway). The amount of incorporated radioactivity, proportional to DNA synthesis, was measured in a liquid scintillation counter (Wallac, Turku, Finland) and expressed as cpm. The S.I. of each experimental group was calculated using the formula: (average cpm of sample in the presence of peptide-pulsed DCs)/(average cpm of sample in the presence of unpulsed DCs).

Immunoblotting

Total cell extracts from 4–5×105 iDCs and DCs matured with LPS, proTα or proTα(100–109) were extracted as described [67]. Briefly, cells were lysed in NP-40 lysis buffer (1% NP-40, 150 mM NaCl, 50 mM Tris pH 8.0) containing protease inhibitors (Protease Inhibitor Cocktail, Sigma-Aldrich) and lysates were cleared by centrifugation for 10 min at 19,000 g (4°C). The protein content of extracts was determined by the Bradford assay, samples were mixed with reducing Laemmli buffer and equal protein amounts (15–25 μg) were separated by sodium dodecyl sulfate–polyacrylamide gel electrophoresis using 12% (w/v) polyacrylamide gels. Separated proteins were blotted on nitrocellulose membranes and probed with primary antibodies (goat anti-human αRIF/Novus Biologicals, Ltd, Cambridge, UK; rabbit anti-human MyD88 and rabbit anti-human TIRAP/eBioscience; rabbit anti-human GAPDH/Santa Cruz Biotechnology Inc, Santa Cruz, CA, USA) and horseradish peroxidase (HRP)-conjugated secondary antibodies (anti-rabbit-IgG and anti-goat-IgG/Santa Cruz Biotechnology). Immunoblots were developed using an enhanced chemiluminescence reagent kit (Santa Cruz Biotechnology) and quantified by scanning densitometry (Gel Analyzer v.1.0, Biosure, Athens, Greece).

Statistical analysis

Data were analyzed by the Student’s t-test and statistical significance was presumed at significance level of 5% (p < 0.05).

Acknowledgments

We thank Prof. S. Stevanovic for providing the anti-MHC antibodies and Dr. I.F. Voutsas for his assistance in flow cytometry analysis. This research has been co-financed by: the European Union (European Social Fund—ESF) and Greek national funds through the Operational Program “Education and Lifelong Learning” of the National Strategic Reference Framework (NSRF)—Research Funding Program: Heracleitus II. Investing in knowledge society through the European Social Fund (to KI); the Hellenic State Scholarship Foundation (IKY) and the Deutscher Akademischer Austauschdienst (DAAD), IKYDA 165/2010 (to WV and OET); NATO SfP Project 982838 (to WV and OET); the European Union FP7 Capacities grant REGPOT-CT-2011-284460, INsPiRE (to IPT and OET) and the BMBF project GerontoShield, grant no. 0315890F (to GP).
Open Access This article is published under license to BioMed Central Ltd. This is an Open Access article is distributed under the terms of the Creative Commons Attribution License ( https://​creativecommons.​org/​licenses/​by/​2.​0 ), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Competing interests

The authors declare that they have no competing interests.

Authors’ contributions

KI: performed the experiments, analyzed data, carried out statistical analyses and wrote the manuscript. ED: designed, analysed and interpreted flow cytometry data and helped to write the manuscript. ET: participated in immunoblotting data acquisition and analyses. PS: performed sample collection and helped to draft the manuscript. HK: carried out peptide synthesis and purification and helped to draft the manuscript. WV: helped in HLA-typing and to draft the manuscript. IPT: participated in the design of the study and reviewed the manuscript. GP: participated in the design and coordination of the study, helped to draft, reviewed and edited the manuscript. OET: conceived, designed and coordinated the study, drafted and reviewed the manuscript. All authors read and approved the final manuscript.
Anhänge

Electronic supplementary material

Literatur
1.
Zurück zum Zitat Chianese-Bullock KA, Irvin WP, Petroni GR, Murphy C, Smolkin M, Olson WC, Coleman E, Boerner SA, Nail CJ, Neese PY, Yuan A, Hogan KT, Slingluff CL: A multipeptide vaccine is safe and elicits T-cell responses in participants with advanced stage ovarian cancer. J Immunother. 2008, 31: 420-430. 10.1097/CJI.0b013e31816dad10.CrossRefPubMed Chianese-Bullock KA, Irvin WP, Petroni GR, Murphy C, Smolkin M, Olson WC, Coleman E, Boerner SA, Nail CJ, Neese PY, Yuan A, Hogan KT, Slingluff CL: A multipeptide vaccine is safe and elicits T-cell responses in participants with advanced stage ovarian cancer. J Immunother. 2008, 31: 420-430. 10.1097/CJI.0b013e31816dad10.CrossRefPubMed
2.
Zurück zum Zitat Barve M, Bender J, Senzer N, Cunningham C, Greco FA, McCune D, Steis R, Khong H, Richards D, Stephenson J, Ganesa P, Nemunaitis J, Ishioka G, Pappen B, Nemunaitis M, Morse M, Mills B, Maples PB, Sherman J, Nemunaitis JJ: Induction of immune responses and clinical efficacy in a phase II trial of IDM-2101, a 10-epitope cytotoxic T-lymphocyte vaccine, in metastatic non-small-cell lung cancer. J Clin Oncol. 2008, 26: 4418-4425. 10.1200/JCO.2008.16.6462.CrossRefPubMed Barve M, Bender J, Senzer N, Cunningham C, Greco FA, McCune D, Steis R, Khong H, Richards D, Stephenson J, Ganesa P, Nemunaitis J, Ishioka G, Pappen B, Nemunaitis M, Morse M, Mills B, Maples PB, Sherman J, Nemunaitis JJ: Induction of immune responses and clinical efficacy in a phase II trial of IDM-2101, a 10-epitope cytotoxic T-lymphocyte vaccine, in metastatic non-small-cell lung cancer. J Clin Oncol. 2008, 26: 4418-4425. 10.1200/JCO.2008.16.6462.CrossRefPubMed
3.
Zurück zum Zitat Walter S, Weinschenk T, Stenzl A, Zdrojowy R, Pluzanska A, Szczylik C, Staehler M, Brugger W, Dietrich PY, Mendrzyk R, Hilf N, Schoor O, Fritsche J, Mahr A, Maurer D, Vass V, Trautwein C, Lewandrowski P, Flohr C, Pohla H, Stanczak JJ, Bronte V, Mandruzzato S, Biedermann T, Pawelec G, Derhovanessian E, Yamagishi H, Miki T, Hongo F, Takaha N, Hirakawa K, Tanaka H, Stevanovic S, Frisch J, Mayer-Mokler A, Kirner A, Rammensee HG, Reinhardt C, Singh-Jasuja H: Multipeptide immune response to cancer vaccine IMA901 after single-dose cyclophosphamide associates with longer patient survival. Nat Med. 2012, 18: 1254-1261. 10.1038/nm.2883.CrossRefPubMed Walter S, Weinschenk T, Stenzl A, Zdrojowy R, Pluzanska A, Szczylik C, Staehler M, Brugger W, Dietrich PY, Mendrzyk R, Hilf N, Schoor O, Fritsche J, Mahr A, Maurer D, Vass V, Trautwein C, Lewandrowski P, Flohr C, Pohla H, Stanczak JJ, Bronte V, Mandruzzato S, Biedermann T, Pawelec G, Derhovanessian E, Yamagishi H, Miki T, Hongo F, Takaha N, Hirakawa K, Tanaka H, Stevanovic S, Frisch J, Mayer-Mokler A, Kirner A, Rammensee HG, Reinhardt C, Singh-Jasuja H: Multipeptide immune response to cancer vaccine IMA901 after single-dose cyclophosphamide associates with longer patient survival. Nat Med. 2012, 18: 1254-1261. 10.1038/nm.2883.CrossRefPubMed
4.
5.
Zurück zum Zitat Jähnisch H, Füssel S, Kiessling A, Wehner R, Zastrow S, Bachmann M, Rieber EP, Wirth MP, Schmitz M: Dendritic cell-based immunotherapy for prostate cancer. Clin Dev Immunol. 2010, 2010: 517493-PubMedCentralCrossRefPubMed Jähnisch H, Füssel S, Kiessling A, Wehner R, Zastrow S, Bachmann M, Rieber EP, Wirth MP, Schmitz M: Dendritic cell-based immunotherapy for prostate cancer. Clin Dev Immunol. 2010, 2010: 517493-PubMedCentralCrossRefPubMed
6.
Zurück zum Zitat Palucka AK, Ueno H, Fay J, Banchereau J: Dendritic cells: a critical player in cancer therapy?. J Immunother. 2008, 31: 793-805. 10.1097/CJI.0b013e31818403bc.PubMedCentralCrossRefPubMed Palucka AK, Ueno H, Fay J, Banchereau J: Dendritic cells: a critical player in cancer therapy?. J Immunother. 2008, 31: 793-805. 10.1097/CJI.0b013e31818403bc.PubMedCentralCrossRefPubMed
7.
Zurück zum Zitat Palucka K, Ueno H, Fay J, Banchereau J: Dendritic cells and immunity against cancer. J Intern Med. 2011, 269: 64-73. 10.1111/j.1365-2796.2010.02317.x.PubMedCentralCrossRefPubMed Palucka K, Ueno H, Fay J, Banchereau J: Dendritic cells and immunity against cancer. J Intern Med. 2011, 269: 64-73. 10.1111/j.1365-2796.2010.02317.x.PubMedCentralCrossRefPubMed
8.
Zurück zum Zitat Schreibelt G, Benitez-Ribas D, Schuurhuis D, Lambeck AJ, van Hout-Kuijer M, Schaft N, Punt CJ, Figdor CG, Adema GJ, de Vries IJ: Commonly used prophylactic vaccines as an alternative for synthetically produced TLR ligands to mature monocyte-derived dendritic cells. Blood. 2010, 116: 564-574. 10.1182/blood-2009-11-251884.CrossRefPubMed Schreibelt G, Benitez-Ribas D, Schuurhuis D, Lambeck AJ, van Hout-Kuijer M, Schaft N, Punt CJ, Figdor CG, Adema GJ, de Vries IJ: Commonly used prophylactic vaccines as an alternative for synthetically produced TLR ligands to mature monocyte-derived dendritic cells. Blood. 2010, 116: 564-574. 10.1182/blood-2009-11-251884.CrossRefPubMed
9.
Zurück zum Zitat Pinzon-Charry A, Maxwell T, López JA: Dendritic cell dysfunction in cancer: a mechanism for immunosuppression. Immunol Cell Biol. 2005, 83: 451-461. 10.1111/j.1440-1711.2005.01371.x.CrossRefPubMed Pinzon-Charry A, Maxwell T, López JA: Dendritic cell dysfunction in cancer: a mechanism for immunosuppression. Immunol Cell Biol. 2005, 83: 451-461. 10.1111/j.1440-1711.2005.01371.x.CrossRefPubMed
10.
Zurück zum Zitat Schilling B, Harasymczuk M, Schuler P, Egan J, Ferrone S, Whiteside TL: IRX-2, a novel immunotherapeutic, enhances functions of human dendritic cells. PLoS One. 2013, 8: e47234-10.1371/journal.pone.0047234.PubMedCentralCrossRefPubMed Schilling B, Harasymczuk M, Schuler P, Egan J, Ferrone S, Whiteside TL: IRX-2, a novel immunotherapeutic, enhances functions of human dendritic cells. PLoS One. 2013, 8: e47234-10.1371/journal.pone.0047234.PubMedCentralCrossRefPubMed
11.
Zurück zum Zitat Napoletano C, Zizzari IG, Rughetti A, Rahimi H, Irimura T, Clausen H, Wandall HH, Belleudi F, Bellati F, Pierelli L, Frati L, Nuti M: Targeting of macrophage galactose-type C-type lectin (MGL) induces DC signaling and activation. Eur J Immunol. 2012, 42: 936-945. 10.1002/eji.201142086.CrossRefPubMed Napoletano C, Zizzari IG, Rughetti A, Rahimi H, Irimura T, Clausen H, Wandall HH, Belleudi F, Bellati F, Pierelli L, Frati L, Nuti M: Targeting of macrophage galactose-type C-type lectin (MGL) induces DC signaling and activation. Eur J Immunol. 2012, 42: 936-945. 10.1002/eji.201142086.CrossRefPubMed
12.
Zurück zum Zitat Vacchelli E, Galluzzi L, Eggermont A, Fridman WH, Galon J, Sautès-Fridman C, Tartour E, Zitvogel L, Kroemer G: Trial watch: FDA-approved Toll-like receptor agonists for cancer therapy. Oncoimmunology. 2012, 1: 894-907. 10.4161/onci.20931.PubMedCentralCrossRefPubMed Vacchelli E, Galluzzi L, Eggermont A, Fridman WH, Galon J, Sautès-Fridman C, Tartour E, Zitvogel L, Kroemer G: Trial watch: FDA-approved Toll-like receptor agonists for cancer therapy. Oncoimmunology. 2012, 1: 894-907. 10.4161/onci.20931.PubMedCentralCrossRefPubMed
13.
Zurück zum Zitat Karbach J, Gnjatic S, Bender A, Neumann A, Weidmann E, Yuan J, Ferrara CA, Hoffmann E, Old LJ, Altorki NK, Jäger E: Tumor-reactive CD8+ T-cell responses after vaccination with NY-ESO-1 peptide, CpG 7909 and Montanide ISA-51: association with survival. Int J Cancer. 2010, 126: 909-918.PubMed Karbach J, Gnjatic S, Bender A, Neumann A, Weidmann E, Yuan J, Ferrara CA, Hoffmann E, Old LJ, Altorki NK, Jäger E: Tumor-reactive CD8+ T-cell responses after vaccination with NY-ESO-1 peptide, CpG 7909 and Montanide ISA-51: association with survival. Int J Cancer. 2010, 126: 909-918.PubMed
14.
Zurück zum Zitat Wieckowski E, Chatta GS, Mailliard RM, Gooding W, Palucka K, Banchereau J, Kalinski P: Type-1 polarized dendritic cells loaded with apoptotic prostate cancer cells are potent inducers of CD8(+) T cells against prostate cancer cells and defined prostate cancer-specific epitopes. Prostate. 2011, 71: 125-133. 10.1002/pros.21228.PubMedCentralCrossRefPubMed Wieckowski E, Chatta GS, Mailliard RM, Gooding W, Palucka K, Banchereau J, Kalinski P: Type-1 polarized dendritic cells loaded with apoptotic prostate cancer cells are potent inducers of CD8(+) T cells against prostate cancer cells and defined prostate cancer-specific epitopes. Prostate. 2011, 71: 125-133. 10.1002/pros.21228.PubMedCentralCrossRefPubMed
15.
Zurück zum Zitat Feyerabend S, Stevanovic S, Gouttefangeas C, Wernet D, Hennenlotter J, Bedke J, Dietz K, Pascolo S, Kuczyk M, Rammensee HG, Stenzl A: Novel multi-peptide vaccination in Hla-A2+ hormone sensitive patients with biochemical relapse of prostate cancer. Prostate. 2009, 69: 917-927. 10.1002/pros.20941.CrossRefPubMed Feyerabend S, Stevanovic S, Gouttefangeas C, Wernet D, Hennenlotter J, Bedke J, Dietz K, Pascolo S, Kuczyk M, Rammensee HG, Stenzl A: Novel multi-peptide vaccination in Hla-A2+ hormone sensitive patients with biochemical relapse of prostate cancer. Prostate. 2009, 69: 917-927. 10.1002/pros.20941.CrossRefPubMed
16.
Zurück zum Zitat Cluff CW: Monophosphoryl lipid A (MPL) as an adjuvant for anti-cancer vaccines: clinical results. Adv Exp Med Biol. 2010, 667: 111-123.CrossRefPubMed Cluff CW: Monophosphoryl lipid A (MPL) as an adjuvant for anti-cancer vaccines: clinical results. Adv Exp Med Biol. 2010, 667: 111-123.CrossRefPubMed
17.
Zurück zum Zitat Butler GS, Overall CM: Proteomic identification of multitasking proteins in unexpected locations complicates drug targeting. Nat Rev Drug Discov. 2009, 8: 935-948. 10.1038/nrd2945.CrossRefPubMed Butler GS, Overall CM: Proteomic identification of multitasking proteins in unexpected locations complicates drug targeting. Nat Rev Drug Discov. 2009, 8: 935-948. 10.1038/nrd2945.CrossRefPubMed
18.
Zurück zum Zitat Ioannou K, Samara P, Livaniou E, Derhovanessian E, Tsitsilonis OE: Prothymosin alpha: a ubiquitous polypeptide with potential use in cancer diagnosis and therapy. Cancer Immunol Immunother. 2012, 61: 599-614. 10.1007/s00262-012-1222-8.CrossRefPubMed Ioannou K, Samara P, Livaniou E, Derhovanessian E, Tsitsilonis OE: Prothymosin alpha: a ubiquitous polypeptide with potential use in cancer diagnosis and therapy. Cancer Immunol Immunother. 2012, 61: 599-614. 10.1007/s00262-012-1222-8.CrossRefPubMed
19.
Zurück zum Zitat Skopeliti M, Kratzer U, Altenberend F, Panayotou G, Kalbacher H, Stevanovic S, Voelter W, Tsitsilonis OE: Proteomic exploitation on prothymosin alpha-induced mononuclear cell activation. Proteomics. 2007, 7: 1814-1824. 10.1002/pmic.200600870.CrossRefPubMed Skopeliti M, Kratzer U, Altenberend F, Panayotou G, Kalbacher H, Stevanovic S, Voelter W, Tsitsilonis OE: Proteomic exploitation on prothymosin alpha-induced mononuclear cell activation. Proteomics. 2007, 7: 1814-1824. 10.1002/pmic.200600870.CrossRefPubMed
20.
Zurück zum Zitat Mosoian A, Teixeira A, Burns CS, Sander LE, Gusella GL, He C, Blander JM, Klotman P, Klotman ME: Prothymosin-alpha inhibits HIV-1 via Toll-like receptor 4-mediated type I interferon induction. Proc Natl Acad Sci USA. 2010, 107: 10178-10183. 10.1073/pnas.0914870107.PubMedCentralCrossRefPubMed Mosoian A, Teixeira A, Burns CS, Sander LE, Gusella GL, He C, Blander JM, Klotman P, Klotman ME: Prothymosin-alpha inhibits HIV-1 via Toll-like receptor 4-mediated type I interferon induction. Proc Natl Acad Sci USA. 2010, 107: 10178-10183. 10.1073/pnas.0914870107.PubMedCentralCrossRefPubMed
21.
Zurück zum Zitat Skopeliti M, Voutsas IF, Klimentzou P, Tsiatas ML, Beck A, Bamias A, Moraki M, Livaniou E, Neagu M, Voelter W, Tsitsilonis OE: The immunologically active site of prothymosin alpha is located at the carboxy-terminus of the polypeptide. Evaluation of its in vitro effects in cancer patients. Cancer Immunol Immunother. 2006, 55: 1247-1257. 10.1007/s00262-005-0108-4.CrossRefPubMed Skopeliti M, Voutsas IF, Klimentzou P, Tsiatas ML, Beck A, Bamias A, Moraki M, Livaniou E, Neagu M, Voelter W, Tsitsilonis OE: The immunologically active site of prothymosin alpha is located at the carboxy-terminus of the polypeptide. Evaluation of its in vitro effects in cancer patients. Cancer Immunol Immunother. 2006, 55: 1247-1257. 10.1007/s00262-005-0108-4.CrossRefPubMed
22.
Zurück zum Zitat Baxevanis CN, Thanos D, Reclos GJ, Anastasopoulos E, Tsokos GC, Papamatheakis J, Papamichail M: Prothymosin alpha enhances human and murine MHC class II surface antigen expression and messenger RNA accumulation. J Immunol. 1992, 148: 1979-1984.PubMed Baxevanis CN, Thanos D, Reclos GJ, Anastasopoulos E, Tsokos GC, Papamatheakis J, Papamichail M: Prothymosin alpha enhances human and murine MHC class II surface antigen expression and messenger RNA accumulation. J Immunol. 1992, 148: 1979-1984.PubMed
23.
Zurück zum Zitat Skopeliti M, Iconomidou VA, Derhovanessian E, Pawelec G, Voelter W, Kalbacher H, Hamodrakas SJ, Tsitsilonis OE: Prothymosin alpha immunoactive carboxyl-terminal peptide TKKQKTDEDD stimulates lymphocyte reactions, induces dendritic cell maturation and adopts a beta-sheet conformation in a sequence-specific manner. Mol Immunol. 2009, 46: 784-792. 10.1016/j.molimm.2008.09.014.CrossRefPubMed Skopeliti M, Iconomidou VA, Derhovanessian E, Pawelec G, Voelter W, Kalbacher H, Hamodrakas SJ, Tsitsilonis OE: Prothymosin alpha immunoactive carboxyl-terminal peptide TKKQKTDEDD stimulates lymphocyte reactions, induces dendritic cell maturation and adopts a beta-sheet conformation in a sequence-specific manner. Mol Immunol. 2009, 46: 784-792. 10.1016/j.molimm.2008.09.014.CrossRefPubMed
24.
Zurück zum Zitat Kapsenberg ML: Dendritic-cell control of pathogen-driven T-cell polarization. Nat Rev Immunol. 2003, 3: 984-993. 10.1038/nri1246.CrossRefPubMed Kapsenberg ML: Dendritic-cell control of pathogen-driven T-cell polarization. Nat Rev Immunol. 2003, 3: 984-993. 10.1038/nri1246.CrossRefPubMed
25.
Zurück zum Zitat Hovden AO, Karlsen M, Jonsson R, Appel S: The bacterial preparation OK432 induces IL-12p70 secretion in human dendritic cells in a TLR3 dependent manner. PLoS One. 2012, 7: e31217-10.1371/journal.pone.0031217.PubMedCentralCrossRefPubMed Hovden AO, Karlsen M, Jonsson R, Appel S: The bacterial preparation OK432 induces IL-12p70 secretion in human dendritic cells in a TLR3 dependent manner. PLoS One. 2012, 7: e31217-10.1371/journal.pone.0031217.PubMedCentralCrossRefPubMed
26.
Zurück zum Zitat Bevan MJ: Helping the CD8(+) T-cell response. Nat Rev Immunol. 2004, 4: 595-602. 10.1038/nri1413.CrossRefPubMed Bevan MJ: Helping the CD8(+) T-cell response. Nat Rev Immunol. 2004, 4: 595-602. 10.1038/nri1413.CrossRefPubMed
27.
Zurück zum Zitat Rubio V, Stuge TB, Singh N, Betts MR, Weber JS, Roederer M, Lee PP: Ex vivo identification, isolation and analysis of tumor-cytolytic T cells. Nat Med. 2003, 9: 1377-1382. 10.1038/nm942.CrossRefPubMed Rubio V, Stuge TB, Singh N, Betts MR, Weber JS, Roederer M, Lee PP: Ex vivo identification, isolation and analysis of tumor-cytolytic T cells. Nat Med. 2003, 9: 1377-1382. 10.1038/nm942.CrossRefPubMed
28.
Zurück zum Zitat Darrah PA, Patel DT, De Luca PM, Lindsay RW, Davey DF, Flynn BJ, Hoff ST, Andersen P, Reed SG, Morris SL, Roederer M, Seder RA: Multifunctional TH1 cells define a correlate of vaccine-mediated protection against Leishmania major. Nat Med. 2007, 13: 843-850. 10.1038/nm1592.CrossRefPubMed Darrah PA, Patel DT, De Luca PM, Lindsay RW, Davey DF, Flynn BJ, Hoff ST, Andersen P, Reed SG, Morris SL, Roederer M, Seder RA: Multifunctional TH1 cells define a correlate of vaccine-mediated protection against Leishmania major. Nat Med. 2007, 13: 843-850. 10.1038/nm1592.CrossRefPubMed
29.
Zurück zum Zitat Precopio ML, Betts MR, Parrino J, Price DA, Gostick E, Ambrozak DR, Asher TE, Douek DC, Harari A, Pantaleo G, Bailer R, Graham BS, Roederer M, Koup RA: Immunization with vaccinia virus induces polyfunctional and phenotypically distinctive CD8(+) T cell responses. J Exp Med. 2007, 204: 1405-1416. 10.1084/jem.20062363.PubMedCentralCrossRefPubMed Precopio ML, Betts MR, Parrino J, Price DA, Gostick E, Ambrozak DR, Asher TE, Douek DC, Harari A, Pantaleo G, Bailer R, Graham BS, Roederer M, Koup RA: Immunization with vaccinia virus induces polyfunctional and phenotypically distinctive CD8(+) T cell responses. J Exp Med. 2007, 204: 1405-1416. 10.1084/jem.20062363.PubMedCentralCrossRefPubMed
30.
Zurück zum Zitat Yamamoto M, Akira S: Lipid A receptor TLR4-mediated signaling pathways. Adv Exp Med Biol. 2010, 667: 59-68.CrossRefPubMed Yamamoto M, Akira S: Lipid A receptor TLR4-mediated signaling pathways. Adv Exp Med Biol. 2010, 667: 59-68.CrossRefPubMed
31.
Zurück zum Zitat Zanoni I, Ostuni R, Marek LR, Barresi S, Barbalat R, Barton GM, Granucci F, Kagan JC: CD14 controls the LPS-induced endocytosis of Toll-like receptor 4. Cell. 2011, 147: 868-880. 10.1016/j.cell.2011.09.051.PubMedCentralCrossRefPubMed Zanoni I, Ostuni R, Marek LR, Barresi S, Barbalat R, Barton GM, Granucci F, Kagan JC: CD14 controls the LPS-induced endocytosis of Toll-like receptor 4. Cell. 2011, 147: 868-880. 10.1016/j.cell.2011.09.051.PubMedCentralCrossRefPubMed
32.
Zurück zum Zitat Nagata A, Takezako N, Tamemoto H, Ohto-Ozaki H, Ohta S, Tominaga S, Yanagisawa K: Soluble ST2 protein inhibits LPS stimulation on monocyte-derived dendritic cells. Cell Mol Immunol. 2012, 9: 399-409. 10.1038/cmi.2012.29.PubMedCentralCrossRefPubMed Nagata A, Takezako N, Tamemoto H, Ohto-Ozaki H, Ohta S, Tominaga S, Yanagisawa K: Soluble ST2 protein inhibits LPS stimulation on monocyte-derived dendritic cells. Cell Mol Immunol. 2012, 9: 399-409. 10.1038/cmi.2012.29.PubMedCentralCrossRefPubMed
33.
Zurück zum Zitat Takeda K, Akira S: TLR signaling pathways. Semin Immunol. 2004, 16: 3-9. 10.1016/j.smim.2003.10.003.CrossRefPubMed Takeda K, Akira S: TLR signaling pathways. Semin Immunol. 2004, 16: 3-9. 10.1016/j.smim.2003.10.003.CrossRefPubMed
34.
Zurück zum Zitat Navabi H, Jasani B, Reece A, Clayton A, Tabi Z, Donninger C, Mason M, Adams M: A clinical grade poly I:C-analogue (Ampligen) promotes optimal DC maturation and Th1-type T cell responses of healthy donors and cancer patients in vitro. Vaccine. 2009, 27: 107-115. 10.1016/j.vaccine.2008.10.024.CrossRefPubMed Navabi H, Jasani B, Reece A, Clayton A, Tabi Z, Donninger C, Mason M, Adams M: A clinical grade poly I:C-analogue (Ampligen) promotes optimal DC maturation and Th1-type T cell responses of healthy donors and cancer patients in vitro. Vaccine. 2009, 27: 107-115. 10.1016/j.vaccine.2008.10.024.CrossRefPubMed
35.
Zurück zum Zitat Kalinski P, Edington H, Zeh HJ, Okada H, Butterfield LH, Kirkwood JM, Bartlett DL: Dendritic cells in cancer immunotherapy: vaccines or autologous transplants?. Immunol Res. 2011, 50: 235-247. 10.1007/s12026-011-8224-z.PubMedCentralCrossRefPubMed Kalinski P, Edington H, Zeh HJ, Okada H, Butterfield LH, Kirkwood JM, Bartlett DL: Dendritic cells in cancer immunotherapy: vaccines or autologous transplants?. Immunol Res. 2011, 50: 235-247. 10.1007/s12026-011-8224-z.PubMedCentralCrossRefPubMed
36.
Zurück zum Zitat Salazar LG, Fikes J, Southwood S, Ishioka G, Knutson KL, Gooley TA, Schiffman K, Disis ML: Immunization of cancer patients with HER-2/neu-derived peptides demonstrating high-affinity binding to multiple class II alleles. Clin Cancer Res. 2003, 9: 5559-5565.PubMed Salazar LG, Fikes J, Southwood S, Ishioka G, Knutson KL, Gooley TA, Schiffman K, Disis ML: Immunization of cancer patients with HER-2/neu-derived peptides demonstrating high-affinity binding to multiple class II alleles. Clin Cancer Res. 2003, 9: 5559-5565.PubMed
37.
Zurück zum Zitat Bernhard H, Salazar L, Schiffman K, Smorlesi A, Schmidt B, Knutson KL, Disis ML: Vaccination against the HER-2/neu oncogenic protein. Endocr Relat Cancer. 2002, 9: 33-44. 10.1677/erc.0.0090033.CrossRefPubMed Bernhard H, Salazar L, Schiffman K, Smorlesi A, Schmidt B, Knutson KL, Disis ML: Vaccination against the HER-2/neu oncogenic protein. Endocr Relat Cancer. 2002, 9: 33-44. 10.1677/erc.0.0090033.CrossRefPubMed
38.
Zurück zum Zitat Green AM, Difazio R, Flynn JL: IFN-γ from CD4 T cells is essential for host survival and enhances CD8 T cell function during Mycobacterium tuberculosis infection. J Immunol. 2013, 190: 270-277. 10.4049/jimmunol.1200061.PubMedCentralCrossRefPubMed Green AM, Difazio R, Flynn JL: IFN-γ from CD4 T cells is essential for host survival and enhances CD8 T cell function during Mycobacterium tuberculosis infection. J Immunol. 2013, 190: 270-277. 10.4049/jimmunol.1200061.PubMedCentralCrossRefPubMed
39.
Zurück zum Zitat Almeida JR, Price DA, Papagno L, Arkoub ZA, Sauce D, Bornstein E, Asher TE, Samri A, Schnuriger A, Theodorou I, Costagliola D, Rouzioux C, Agut H, Marcelin AG, Douek D, Autran B, Appay V: Superior control of HIV-1 replication by CD8+ T cells is reflected by their avidity, polyfunctionality, and clonal turnover. J Exp Med. 2007, 204: 2473-2485. 10.1084/jem.20070784.PubMedCentralCrossRefPubMed Almeida JR, Price DA, Papagno L, Arkoub ZA, Sauce D, Bornstein E, Asher TE, Samri A, Schnuriger A, Theodorou I, Costagliola D, Rouzioux C, Agut H, Marcelin AG, Douek D, Autran B, Appay V: Superior control of HIV-1 replication by CD8+ T cells is reflected by their avidity, polyfunctionality, and clonal turnover. J Exp Med. 2007, 204: 2473-2485. 10.1084/jem.20070784.PubMedCentralCrossRefPubMed
40.
Zurück zum Zitat Lichtenegger FS, Mueller K, Otte B, Beck B, Hiddemann W, Schendel DJ, Subklewe M: CD86 and IL-12p70 are key players for T helper 1 polarization and natural killer cell activation by Toll-like receptor-induced dendritic cells. PLoS One. 2012, 7: e44266-10.1371/journal.pone.0044266.PubMedCentralCrossRefPubMed Lichtenegger FS, Mueller K, Otte B, Beck B, Hiddemann W, Schendel DJ, Subklewe M: CD86 and IL-12p70 are key players for T helper 1 polarization and natural killer cell activation by Toll-like receptor-induced dendritic cells. PLoS One. 2012, 7: e44266-10.1371/journal.pone.0044266.PubMedCentralCrossRefPubMed
41.
Zurück zum Zitat Jung ID, Jeong SK, Lee CM, Noh KT, Heo DR, Shin YK, Yun CH, Koh WJ, Akira S, Whang J, Kim HJ, Park WS, Shin SJ, Park YM: Enhanced efficacy of therapeutic cancer vaccines produced by co-treatment with Mycobacterium tuberculosis heparin-binding hemagglutinin, a novel TLR4 agonist. Cancer Res. 2011, 71: 2858-2870. 10.1158/0008-5472.CAN-10-3487.CrossRefPubMed Jung ID, Jeong SK, Lee CM, Noh KT, Heo DR, Shin YK, Yun CH, Koh WJ, Akira S, Whang J, Kim HJ, Park WS, Shin SJ, Park YM: Enhanced efficacy of therapeutic cancer vaccines produced by co-treatment with Mycobacterium tuberculosis heparin-binding hemagglutinin, a novel TLR4 agonist. Cancer Res. 2011, 71: 2858-2870. 10.1158/0008-5472.CAN-10-3487.CrossRefPubMed
42.
Zurück zum Zitat Baxevanis CN, Voutsas IF, Tsitsilonis OE: Toll-like receptor agonists: current status and future perspective on their utility as adjuvants in improving anticancer vaccination strategies. Immunotherapy. 2013, 5: 497-511. 10.2217/imt.13.24.CrossRefPubMed Baxevanis CN, Voutsas IF, Tsitsilonis OE: Toll-like receptor agonists: current status and future perspective on their utility as adjuvants in improving anticancer vaccination strategies. Immunotherapy. 2013, 5: 497-511. 10.2217/imt.13.24.CrossRefPubMed
43.
Zurück zum Zitat Bianchi ME: DAMPs, PAMPs and alarmins: all we need to know about danger. J Leukoc Biol. 2007, 81: 1-5.CrossRefPubMed Bianchi ME: DAMPs, PAMPs and alarmins: all we need to know about danger. J Leukoc Biol. 2007, 81: 1-5.CrossRefPubMed
44.
Zurück zum Zitat Messmer D, Yang H, Telusma G, Knoll F, Li J, Messmer B, Tracey KJ, Chiorazzi N: High mobility group box protein 1: an endogenous signal for dendritic cell maturation and Th1 polarization. J Immunol. 2004, 173: 307-313.CrossRefPubMed Messmer D, Yang H, Telusma G, Knoll F, Li J, Messmer B, Tracey KJ, Chiorazzi N: High mobility group box protein 1: an endogenous signal for dendritic cell maturation and Th1 polarization. J Immunol. 2004, 173: 307-313.CrossRefPubMed
45.
Zurück zum Zitat Saenz R, Souza Cda S, Huang CT, Larsson M, Esener S, Messmer D: HMGB1-derived peptide acts as adjuvant inducing immune responses to peptide and protein antigen. Vaccine. 2010, 28: 7556-7562. 10.1016/j.vaccine.2010.08.054.PubMedCentralCrossRefPubMed Saenz R, Souza Cda S, Huang CT, Larsson M, Esener S, Messmer D: HMGB1-derived peptide acts as adjuvant inducing immune responses to peptide and protein antigen. Vaccine. 2010, 28: 7556-7562. 10.1016/j.vaccine.2010.08.054.PubMedCentralCrossRefPubMed
46.
Zurück zum Zitat Wang Y, Kelly CG, Singh M, McGowan EG, Carrara AS, Bergmeier LA, Lehner T: Stimulation of Th1-polarizing cytokines, C-C chemokines, maturation of dendritic cells, and adjuvant function by the peptide binding fragment of heat shock protein 70. J Immunol. 2002, 169: 2422-2429.CrossRefPubMed Wang Y, Kelly CG, Singh M, McGowan EG, Carrara AS, Bergmeier LA, Lehner T: Stimulation of Th1-polarizing cytokines, C-C chemokines, maturation of dendritic cells, and adjuvant function by the peptide binding fragment of heat shock protein 70. J Immunol. 2002, 169: 2422-2429.CrossRefPubMed
47.
Zurück zum Zitat Wu Y, Wan T, Zhou X, Wang B, Yang F, Li N, Chen G, Dai S, Liu S, Zhang M, Cao X: Hsp70-like protein 1 fusion protein enhances induction of carcinoembryonic antigen-specific CD8+ CTL response by dendritic cell vaccine. Cancer Res. 2005, 65: 4947-4954. 10.1158/0008-5472.CAN-04-3912.CrossRefPubMed Wu Y, Wan T, Zhou X, Wang B, Yang F, Li N, Chen G, Dai S, Liu S, Zhang M, Cao X: Hsp70-like protein 1 fusion protein enhances induction of carcinoembryonic antigen-specific CD8+ CTL response by dendritic cell vaccine. Cancer Res. 2005, 65: 4947-4954. 10.1158/0008-5472.CAN-04-3912.CrossRefPubMed
48.
Zurück zum Zitat Tewary P, Yang D, de la Rosa G, Li Y, Finn MW, Krensky AM, Clayberger C, Oppenheim JJ: Granulysin activates antigen-presenting cells through TLR4 and acts as an immune alarmin. Blood. 2010, 116: 3465-3474. 10.1182/blood-2010-03-273953.PubMedCentralCrossRefPubMed Tewary P, Yang D, de la Rosa G, Li Y, Finn MW, Krensky AM, Clayberger C, Oppenheim JJ: Granulysin activates antigen-presenting cells through TLR4 and acts as an immune alarmin. Blood. 2010, 116: 3465-3474. 10.1182/blood-2010-03-273953.PubMedCentralCrossRefPubMed
49.
Zurück zum Zitat Cordero OJ, Sarandeses C, López-Rodríguez JL, Nogueira M: The presence and cytotoxicity of CD16+ CD2- subset from PBL and NK cells in long-term IL-2 cultures enhanced by Prothymosin-alpha. Immunopharmacology. 1995, 29: 215-223. 10.1016/0162-3109(95)00057-Z.CrossRefPubMed Cordero OJ, Sarandeses C, López-Rodríguez JL, Nogueira M: The presence and cytotoxicity of CD16+ CD2- subset from PBL and NK cells in long-term IL-2 cultures enhanced by Prothymosin-alpha. Immunopharmacology. 1995, 29: 215-223. 10.1016/0162-3109(95)00057-Z.CrossRefPubMed
50.
Zurück zum Zitat Eckert K, Grünberg E, Garbin F, Maurer HR: Preclinical studies with prothymosin alpha1 on mononuclear cells from tumor patients. Int J Immunopharmacol. 1997, 19: 493-500. 10.1016/S0192-0561(97)00079-9.CrossRefPubMed Eckert K, Grünberg E, Garbin F, Maurer HR: Preclinical studies with prothymosin alpha1 on mononuclear cells from tumor patients. Int J Immunopharmacol. 1997, 19: 493-500. 10.1016/S0192-0561(97)00079-9.CrossRefPubMed
51.
Zurück zum Zitat Baxevanis CN, Gritzapis AD, Spanakos G, Tsitsilonis OE, Papamichail M: Induction of tumor-specific T lymphocyte responses in vivo by prothymosin alpha. Cancer Immunol Immunother. 1995, 40: 410-418. 10.1007/BF01525392.CrossRefPubMed Baxevanis CN, Gritzapis AD, Spanakos G, Tsitsilonis OE, Papamichail M: Induction of tumor-specific T lymphocyte responses in vivo by prothymosin alpha. Cancer Immunol Immunother. 1995, 40: 410-418. 10.1007/BF01525392.CrossRefPubMed
52.
Zurück zum Zitat Voutsas IF, Baxevanis CN, Gritzapis AD, Missitzis I, Stathopoulos GP, Archodakis G, Banis C, Voelter W, Papamichail M: Synergy between interleukin-2 and prothymosin alpha for the increased generation of cytotoxic T lymphocytes against autologous human carcinomas. Cancer Immunol Immunother. 2000, 49: 449-458. 10.1007/s002620000132.CrossRefPubMed Voutsas IF, Baxevanis CN, Gritzapis AD, Missitzis I, Stathopoulos GP, Archodakis G, Banis C, Voelter W, Papamichail M: Synergy between interleukin-2 and prothymosin alpha for the increased generation of cytotoxic T lymphocytes against autologous human carcinomas. Cancer Immunol Immunother. 2000, 49: 449-458. 10.1007/s002620000132.CrossRefPubMed
53.
Zurück zum Zitat Telusma G, Datta S, Mihajlov I, Ma W, Li J, Yang H, Newman W, Messmer BT, Minev B, Schmidt-Wolf IG, Tracey KJ, Chiorazzi N, Messmer D: Dendritic cell activating peptides induce distinct cytokine profiles. Int Immunol. 2006, 18: 1563-1573. 10.1093/intimm/dxl089.CrossRefPubMed Telusma G, Datta S, Mihajlov I, Ma W, Li J, Yang H, Newman W, Messmer BT, Minev B, Schmidt-Wolf IG, Tracey KJ, Chiorazzi N, Messmer D: Dendritic cell activating peptides induce distinct cytokine profiles. Int Immunol. 2006, 18: 1563-1573. 10.1093/intimm/dxl089.CrossRefPubMed
54.
Zurück zum Zitat Park JS, Svetkauskaite D, He Q, Kim JY, Strassheim D, Ishizaka A, Abraham E: Involvement of toll-like receptors 2 and 4 in cellular activation by high mobility group box 1 protein. J Biol Chem. 2004, 279: 7370-7377.CrossRefPubMed Park JS, Svetkauskaite D, He Q, Kim JY, Strassheim D, Ishizaka A, Abraham E: Involvement of toll-like receptors 2 and 4 in cellular activation by high mobility group box 1 protein. J Biol Chem. 2004, 279: 7370-7377.CrossRefPubMed
55.
Zurück zum Zitat Park JS, Gamboni-Robertson F, He Q, Svetkauskaite D, Kim JY, Strassheim D, Sohn JW, Yamada S, Maruyama I, Banerjee A, Ishizaka A, Abraham E: High mobility group box 1 protein interacts with multiple Toll-like receptors. Am J Physiol Cell Physiol. 2006, 290: C917-924.CrossRefPubMed Park JS, Gamboni-Robertson F, He Q, Svetkauskaite D, Kim JY, Strassheim D, Sohn JW, Yamada S, Maruyama I, Banerjee A, Ishizaka A, Abraham E: High mobility group box 1 protein interacts with multiple Toll-like receptors. Am J Physiol Cell Physiol. 2006, 290: C917-924.CrossRefPubMed
56.
Zurück zum Zitat Yu M, Wang H, Ding A, Golenbock DT, Latz E, Czura CJ, Fenton MJ, Tracey KJ, Yang H: HMGB1 signals through toll-like receptor (TLR) 4 and TLR2. Shock. 2006, 26: 174-179. 10.1097/01.shk.0000225404.51320.82.CrossRefPubMed Yu M, Wang H, Ding A, Golenbock DT, Latz E, Czura CJ, Fenton MJ, Tracey KJ, Yang H: HMGB1 signals through toll-like receptor (TLR) 4 and TLR2. Shock. 2006, 26: 174-179. 10.1097/01.shk.0000225404.51320.82.CrossRefPubMed
57.
Zurück zum Zitat Roelofs MF, Boelens WC, Joosten LA, Abdollahi-Roodsaz S, Geurts J, Wunderink LU, Schreurs BW, van den Berg WB, Radstake TR: Identification of small heat shock protein B8 (HSP22) as a novel TLR4 ligand and potential involvement in the pathogenesis of rheumatoid arthritis. J Immunol. 2006, 176: 7021-7027.CrossRefPubMed Roelofs MF, Boelens WC, Joosten LA, Abdollahi-Roodsaz S, Geurts J, Wunderink LU, Schreurs BW, van den Berg WB, Radstake TR: Identification of small heat shock protein B8 (HSP22) as a novel TLR4 ligand and potential involvement in the pathogenesis of rheumatoid arthritis. J Immunol. 2006, 176: 7021-7027.CrossRefPubMed
58.
Zurück zum Zitat Vabulas RM, Ahmad-Nejad P, da Costa C, Miethke T, Kirschning CJ, Häcker H, Wagner HL: 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. 2001, 276: 31332-31339. 10.1074/jbc.M103217200.CrossRefPubMed Vabulas RM, Ahmad-Nejad P, da Costa C, Miethke T, Kirschning CJ, Häcker H, Wagner HL: 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. 2001, 276: 31332-31339. 10.1074/jbc.M103217200.CrossRefPubMed
59.
Zurück zum Zitat Vabulas RM, Ahmad-Nejad P, Ghose S, Kirschning CJ, Issels RD, Wagner H: HSP70 as endogenous stimulus of the Toll/interleukin-1 receptor signal pathway. J Biol Chem. 2002, 277: 15107-15112. 10.1074/jbc.M111204200.CrossRefPubMed Vabulas RM, Ahmad-Nejad P, Ghose S, Kirschning CJ, Issels RD, Wagner H: HSP70 as endogenous stimulus of the Toll/interleukin-1 receptor signal pathway. J Biol Chem. 2002, 277: 15107-15112. 10.1074/jbc.M111204200.CrossRefPubMed
60.
Zurück zum Zitat Jiang XX, Zhang Y, Liu B, Zhang SX, Wu Y, Yu XD, Mao N: Human mesenchymal stem cells inhibit differentiation and function of monocyte-derived dendritic cells. Blood. 2005, 105: 4120-4126. 10.1182/blood-2004-02-0586.CrossRefPubMed Jiang XX, Zhang Y, Liu B, Zhang SX, Wu Y, Yu XD, Mao N: Human mesenchymal stem cells inhibit differentiation and function of monocyte-derived dendritic cells. Blood. 2005, 105: 4120-4126. 10.1182/blood-2004-02-0586.CrossRefPubMed
61.
Zurück zum Zitat Jarnjak-Jankovic S, Hammerstad H, Saebøe-Larssen S, Kvalheim G, Gaudernack G: A full scale comparative study of methods for generation of functional dendritic cells for use as cancer vaccines. BMC Cancer. 2007, 7: 119-10.1186/1471-2407-7-119.PubMedCentralCrossRefPubMed Jarnjak-Jankovic S, Hammerstad H, Saebøe-Larssen S, Kvalheim G, Gaudernack G: A full scale comparative study of methods for generation of functional dendritic cells for use as cancer vaccines. BMC Cancer. 2007, 7: 119-10.1186/1471-2407-7-119.PubMedCentralCrossRefPubMed
62.
Zurück zum Zitat Zanoni I, Granucci F: Differences in lipopolysaccharide-induced signaling between conventional dendritic cells and macrophages. Immunobiology. 2010, 215: 709-712. 10.1016/j.imbio.2010.05.026.CrossRefPubMed Zanoni I, Granucci F: Differences in lipopolysaccharide-induced signaling between conventional dendritic cells and macrophages. Immunobiology. 2010, 215: 709-712. 10.1016/j.imbio.2010.05.026.CrossRefPubMed
63.
Zurück zum Zitat Bosisio D, Polentarutti N, Sironi M, Bernasconi S, Miyake K, Webb GR, Martin MU, Mantovani A, Muzio M: Stimulation of toll-like receptor 4 expression in human mononuclear phagocytes by interferon-gamma: a molecular basis for priming and synergism with bacterial lipopolysaccharide. Blood. 2002, 99: 3427-3431. 10.1182/blood.V99.9.3427.CrossRefPubMed Bosisio D, Polentarutti N, Sironi M, Bernasconi S, Miyake K, Webb GR, Martin MU, Mantovani A, Muzio M: Stimulation of toll-like receptor 4 expression in human mononuclear phagocytes by interferon-gamma: a molecular basis for priming and synergism with bacterial lipopolysaccharide. Blood. 2002, 99: 3427-3431. 10.1182/blood.V99.9.3427.CrossRefPubMed
64.
Zurück zum Zitat Robbins PF: Tumor associated antigens. Analyzing T cell responses. Edited by: Nagorsen D, Marincola FM. 2005, Netherlands: Springer, 9-42.CrossRef Robbins PF: Tumor associated antigens. Analyzing T cell responses. Edited by: Nagorsen D, Marincola FM. 2005, Netherlands: Springer, 9-42.CrossRef
65.
Zurück zum Zitat Gavalas NG, Tsiatas M, Tsitsilonis O, Politi E, Ioannou K, Ziogas AC, Rodolakis A, Vlahos G, Thomakos N, Haidopoulos D, Terpos E, Antsaklis A, Dimopoulos MA, Bamias A: VEGF directly suppresses activation of T cells from ascites secondary to ovarian cancer via VEGF receptor type 2. Br J Cancer. 2012, 107: 1869-1875. 10.1038/bjc.2012.468.PubMedCentralCrossRefPubMed Gavalas NG, Tsiatas M, Tsitsilonis O, Politi E, Ioannou K, Ziogas AC, Rodolakis A, Vlahos G, Thomakos N, Haidopoulos D, Terpos E, Antsaklis A, Dimopoulos MA, Bamias A: VEGF directly suppresses activation of T cells from ascites secondary to ovarian cancer via VEGF receptor type 2. Br J Cancer. 2012, 107: 1869-1875. 10.1038/bjc.2012.468.PubMedCentralCrossRefPubMed
66.
Zurück zum Zitat Sun Y, Stevanovic S, Song M, Schwantes A, Kirkpatrick CJ, Schadendorf D, Cichutek K: The kinase insert domain-containing receptor is an angiogenesis-associated antigen recognized by human cytotoxic T lymphocytes. Blood. 2006, 107: 1476-1483. 10.1182/blood-2005-05-1912.CrossRefPubMed Sun Y, Stevanovic S, Song M, Schwantes A, Kirkpatrick CJ, Schadendorf D, Cichutek K: The kinase insert domain-containing receptor is an angiogenesis-associated antigen recognized by human cytotoxic T lymphocytes. Blood. 2006, 107: 1476-1483. 10.1182/blood-2005-05-1912.CrossRefPubMed
67.
Zurück zum Zitat Antonelou MH, Kriebardis AG, Stamoulis KE, Trougakos IP, Papassideri IS: Apolipoprotein J/clusterin in human erythrocytes is involved in the molecular process of defected material disposal during vesiculation. PLoS One. 2011, 6: e26033-10.1371/journal.pone.0026033.PubMedCentralCrossRefPubMed Antonelou MH, Kriebardis AG, Stamoulis KE, Trougakos IP, Papassideri IS: Apolipoprotein J/clusterin in human erythrocytes is involved in the molecular process of defected material disposal during vesiculation. PLoS One. 2011, 6: e26033-10.1371/journal.pone.0026033.PubMedCentralCrossRefPubMed
Metadaten
Titel
Prothymosin α and a prothymosin α-derived peptide enhance TH1-type immune responses against defined HER-2/neu epitopes
verfasst von
Kyriaki Ioannou
Evelyna Derhovanessian
Eleni Tsakiri
Pinelopi Samara
Hubert Kalbacher
Wolfgang Voelter
Ioannis P Trougakos
Graham Pawelec
Ourania E Tsitsilonis
Publikationsdatum
01.12.2013
Verlag
BioMed Central
Erschienen in
BMC Immunology / Ausgabe 1/2013
Elektronische ISSN: 1471-2172
DOI
https://doi.org/10.1186/1471-2172-14-43

Weitere Artikel der Ausgabe 1/2013

BMC Immunology 1/2013 Zur Ausgabe

Leitlinien kompakt für die Innere Medizin

Mit medbee Pocketcards sicher entscheiden.

Seit 2022 gehört die medbee GmbH zum Springer Medizin Verlag

Update Innere Medizin

Bestellen Sie unseren Fach-Newsletter und bleiben Sie gut informiert.