Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2014 May 30.
Published in final edited form as: Virus Res. 2007 Nov 26;132(0):1–14. doi: 10.1016/j.virusres.2007.10.005

Adenovirus vector induced Innate Immune responses: Impact upon efficacy and toxicity in gene therapy and vaccine applications

Zachary C Hartman 2,*, Daniel M Appledorn 1,*, Andrea Amalfitano 1,2
PMCID: PMC4039020  NIHMSID: NIHMS42051  PMID: 18036698

Abstract

Extensively characterized, modified, and employed for a variety of purposes, Adenovirus (Ad) vectors are generally regarded as having great potential by many applied virologists who wish to manipulate and use viral biology to achieve beneficial clinical outcomes. Despite widespread functional prominence and utility, (i.e.: Ad based clinical trials have begun to progress to critical Phase III levels, it has recently become apparent that investigations regarding the innate immune response to Ads may reveal not only reasons behind previous failures, but also reveal novel insights that will allow for safer, more efficacious uses of this important gene transfer platform. Insights gained by the exploration of Ad induced innate immune responses will likely be most important to the fields of vaccine development, since Ad based vaccines are highly acknowledged as one of the more promising vaccine platforms in development today.

Adenovirus is currently known to interact with several different extracellular, intracellular, and membrane bound innate immune sensing systems. Past and recent studies involving manipulation of the Ad infectious cycle as well as use of different mutants have shed light on some of the initiation mechanisms underlying Ad induced immune responses. More recent studies using microarray based analyses, genetically modified cell lines and/or mouse mutants, and advanced generation Ad vectors have revealed important new insights into the scope and mechanism of this cellular defensive response. This review is an attempt to synthesize these studies, update Ad biologists to the current knowledge surrounding these increasingly important issues, as well point areas where future research should be directed. It should also serve as a sobering reality to researchers exploring the use of any gene transfer vector, as to the complexities potentially involved when contemplating use of such vectors for human applications.

Keywords: Adenovirus, Gene Therapy, Review, Innate Immunity, Complement, Humoral Immunity, Cellular Immunity, Toll-Like Receptors

Introduction

Adenovirus (Ad) based gene transfer has been an important part of the scientific landscape, contributing to a great number of studies, both of basic virology, as well as more focused and recent uses in gene therapy and vaccine applications. As of this date, more human clinical trials utilize recombinant Ad (rAd) based vectors than any other gene transfer platform. Furthermore, recent progress in utilizing Ad based vectors as a vaccine platform suggests greater uses of this gene transfer workhorse for expanding purposes, such as in HIV-AIDS, cancer immunotherapy approaches, and in vaccination for serious infections such as Ebola, and bird flu. Despite these successes, the scientific and bio-industrial communities have recently recognized that a great lack of knowledge exists regarding the mechanisms by which Ad vectors induce several classes of innate immune responses. Unfortunately this has, in some isolated cases, resulted in serious consequences. Clearly, a better understanding of the Ad induced innate immune responses will not only foster safer utilization of this platform, but such knowledge will also greatly improve future applications of this increasingly useful gene transfer platform.

Since very early in the evolutionary history of multi-cellular organisms, cellular and viral parasites have emerged to pose a significant challenge to a complex organism’s survival (Hoffmann et al., 1999). In mammals, this challenge was met by the primordial co-evolution of defensive networks that include both the innate and adaptive immune systems. The principal task of these systems was to evolve in a manner that allowed the host to rapidly detect the first encounter with a potential pathogen, coordinate both isolation and neutralization of the invader, and to create a “rapid-response” defense mechanism to be triggered in the event of repeat encounter (Beutler and Rietschel, 2003; Hoffmann et al., 1999; Medzhitov and Janeway, 2002). In light of this evolutionary arms race, the difficulty surrounding the conversion of viruses into vectors for gene transfer should come as little surprise.

The use of viral vectors for gene transfer is a relatively simple idea; insertion of desired genetic material into the viral genome, thereby allowing one to take advantage of the inherent ability of the virus to transduce cells for a desired therapeutic outcome. For the past twenty years, this strategy has been applied to the one of the most extensively studied and widely used viral vector systems, those based upon human adenoviruses (Ads). Named for the adenoid tissue from which it was initially isolated, the genus Adenoviridae encompasses a large family of non-enveloped, double-stranded DNA viruses. From this family, human subgroup C Ads have been the most extensively characterized, with two of its serotypes, 2 and 5, emerging as the viral platforms most commonly used for gene delivery and vaccination purposes. While work involving the genetic manipulation of the virus has been widely reported, much recent research has focused on analyzing Ad vector interactions with the innate immune system. We feel that a clearer understanding of this knowledge will help direct future research efforts attempting to improve the safety and/or more effective use of recombinant Ad-based vectors in widespread clinical applications. This review will therefore examine the innate immune repertoire elicited by Ads, or Ad-based vectors, both in vitro and in vivo, including their interactions with important extracellular and intracellular pathogen sensing mechanisms.

Ad INDUCED CELLULAR DEFENSE RESPONSES IN VITRO

More than 30 years ago, in vitro studies using adenovirus found it to be a potent inducer of ‘interferon’ in chick embryonic fibroblasts, setting the stage for the exploration of innate immune responses to adenovirus in vitro (Tarodi et al., 1977; Ustacelebi and Williams, 1972). Since that time there has been considerable research in this area, and it is now well-appreciated that in vitro Ad infection of cells initiates a broad series of events, inclusive of stress response pathways, metabolic changes, and intrinsic immune responses. The ability to generate high titers of Ad vector has facilitated in vitro analyses of Ad-specific defense responses, as lower Ad titers may activate similar responses, but at levels that may be undetectable with current technologies. As an aside, these considerations should be understood when comparisons are made between gene transfer (viral) vectors, as a “lack” of detection of immune responses could be due to inadequate assay sensitivity. Here we will examine the cellular defensive responses of both ‘non-immune’ cells and more specialized ‘immune’ cell types to Ad infection, highlighting common patterns as well as insights into the mechanism of this response. We will then examine recent work regarding the molecular mechanisms underlying these responses in an examination of the pathogen recognition receptor (PRR) contribution to adenoviral induced innate responses in vitro.

“Non-Immune” Cell Responses to Ad infection in vitro

Ad infection of what many would consider “non-immune” cells has been found to precipitate extremely rapid changes, with p42/MAPK phosphorylation and ERK signaling occurring 10–20 minutes after Ad infection in both HeLa (human cervical epithelial cells) and A549 (human respiratory epithelial cells), the latter of which was shown to be, in part, dependent upon fiber/CAR interactions, and independent of penton-RGD integrin interactions (Tamanini et al., 2006; Tibbles et al., 2002). Similar signaling events occur in Ad infected REC (murine renal epithelium-derived) cells, with p38 and ERK activation occurring 10 minutes post-infection. These early signaling events were directly linked to the subsequent expression of chemokine genes, as pharmacological inhibitors of p38 and ERK directly inhibited Ad triggered IP-10 expression (Liu and Muruve, 2003; Bowen et al., 2002; Bruder and Kovesdi, 1997; Tibbles et al., 2002). In a follow-up study, Liu et al., (2005) showed that within 10 minutes, RGD/integrin interactions resulted in PI3K dependent Akt and NF-κB activation and subsequent IP-10 transcription. Other genes induced by Ad infection included various leukocyte adhesion molecules, such as ICAM-1 and VCAM-1. In human embryonic derived 293 (HEK293) cells, UV-inactivated Ad vectors induce the expression of several chemokines, particularly RANTES, IP-10, and MIP-2 (Liu and Muruve, 2003). Furthermore, HeLa, A549, and the TGP61 mouse insulinoma cell line, show similar innate chemokine activation patterns, with Ad infections inducing expression of IP-10, RANTES, and in human cells, IL-8. In contrast to findings in macrophages, the induction of RANTES and IP-10 in REC epithelial cells and in HeLa cells occurred in the absence of a significant TNF-α and IL-1β induction. Nuclear translocation of NF-κB also figures prominently in Ad induced innate induction of cytokines and chemokines, both in cell culture and in vivo (Borgland et al., 2000; Bowen et al., 2002; Loser et al., 1998; Muruve et al., 1999). These data demonstrate that Ad infection activates broad cellular responses in what many would consider “non-immune cells”, such as epithelial and endothelial cells.

Furthermore, different non-immune cell types respond to Ad infection by the production of different sets of cytokines, although the specific repertoire of cytokines elaborated likely varies between cell types.

In contrast to the approach taken in the aforementioned studies, (where innate immune activation was probed through the investigation of single genes or kinases) several investigations have explored the effects of Ads on non-immune cells using a more global approach, involving microarray based transcriptome analysis. For example, Stilwell and Samulski, (2004) found that first generation recombinant Ads (FG-Ads), as well as ‘empty Ad capsids’ altered a significant portion of the cellular transcriptome. Specifically, 24 hours post-infection (hpi) of IMR-90 human primary lung fibroblasts, FG-Ads caused 3-fold or greater changes in the expression of 4.2% of the genes examined, while ‘empty’ Ad capsids caused expressional changes in 0.7% of the genes examined, at . Despite the fact that the ‘empty Ad capsids’ possessed a somewhat altered morphology, and lacked any encapsulated DNA, (thus no terminal protein), they still elicited a sizable number of transcriptome changes in this non-immune human diploid embryonic fibroblast cell line, constituting a putative “capsid induced” transcriptome signature.

In an investigation of tropism-modified Ads, Volk et al., (2005) found rather striking differences in the global gene transcription profiles induced by Ad transduction of a human melanoma cell line (M21) with 1) native FG-Ads, 2) an Ad5 virus displaying the Ad3 fiber knob domain (Ad5/3) which alters the interaction with the coxsackie-Adenovirus receptor (CAR), and 3) an Ad5 vector displaying a fiber knob with an additional RGD binding domain. All three vectors induced dysregulation of the cellular transcriptome at 30hpi, with the native Ad5 vector inducing the most dysregulation, followed by the Ad5/3 chimera and the RGD fiber knob mutant Ad. The common group of genes whose expression was dysregulated by the different Ads showed little common functionality, although a statistical grouping of over-represented genes that may have had consensus functions was not performed. This study highlights the importance of the adenoviral entry process upon cellular defensive responses elicited after infection, and should serve as an example of the complexities one may encounter when proposing to utilize detargeted, retargeted or alternative serotype Ad vectors.

In a similar comparative analysis, Martina et al., (2007) characterized the transcriptome response induced by infection of liver cells (Huh-7) with wild-type Ad5 (wtAd5), and compared this response to infections of cells with either FG-Ad5 (deleted for E1 and E3 genes) or a helper dependent Ad (HD-Ad). The HD-Ad contained full length stuffer DNA content, but retained only the ITRs of the original Ad genome (Parks et al., 1996). Based strictly upon fold-change analysis, at 4hpi wtAd5 dysregulated the expression of 905 Huh-7 genes, while the FG-Ad5 dysregulated 552 (61% of wt-Ad5), and the HD-Ad dysregulated 300 (33% of wt-Ad5) Huh-7 genes; at 12hpi Ad5 dysregulated the expression of 1649 Huh-7 genes, while the FG-Ad5 dysregulated expression of 527 (32% of wt-Ad5), and the HD-Ad dysregulated expression of 658 (40% of wt-Ad5) Huh-7 genes. Though the modified Ads dysregulated a sizeable number of genes relative to similar infections with wtAd5, only a small proportion of the genes overlapped based on functional or pathway analysis based methods. Importantly, these investigators also noted that although HD-Ad5 and FG-Ad5 both induced similar levels of transcriptome dysregulation (relative to the wild-type Ad5), only ≈8% of these changes were shared between the two. This lack of overlap indirectly suggests that specific DNA sequences contained within the adenoviral vector may impact upon other cellular responses, although future studies will be required to confirm this hypothesis.

Addressing the issue of Ad induction of cellular transcriptome responses over time, Dorn et al., (2005) assayed global gene expression changes at 12, 32, and 48hpi in HeLa cells after infection with wild-type Ad12. This investigation determined that maximal gene induction of interferon-related genes in Ad12 infected HeLa cells occurs at 12hpi, with strong global repression of most other genes occurring at the later time points. Although this study used a wild-type Ad (in contrast to the ablated vectors mentioned previously), it did demonstrate the importance of time as a factor in the course of the overall cellular innate immune response, with the complexities surrounding the transcriptome responses during a wild-type Ad infection being a sub-aspect of the total response.

These analyses began to identify a core of Ad dysregulated genes, although the architecture of the signaling cascade governing Ad innate immune responses (as reflected by global transcriptome responses) remained largely uncharacterized. Our group recently completed a study in Ad vector infected mouse embryonic fibroblasts (MEFs) in an attempt to identify “Ad gene response” networks (Hartman et al., 2007a). At 16hpi, we detected dysregulation of 22% of the total number of gene probes assessable by the array, 75% of which were upregulated at least 2-fold, and 33% that were upregulated 3-fold or higher. Possibly due to the fact that MEFs express a full Toll-Like Receptor (TLR) repertoire, major Ad-affected gene ontology categories included immune response and antigen processing genes. However, cell cycle, hydrolase activity, heat shock and cytokinesis genes were also over-represented. UV inactivation of the Ad or cyclohexamide treatment of the Ad infected MEF cells did not affect the expression of a subset of these genes, strengthening the notion that Ad virions initiate rapid host transcriptome responses independent of Ad transcription, DNA replication and/or protein expression. In many cases, UV-inactivated virus resulted in higher levels of gene upregulation suggesting Ad genome derived transcription products may actually repress capsid (and/or cytoplasmically sensed “Ad DNA”) -initiated responses.

To indirectly identify transcription factors responsible for Ad-induced or suppressed gene networks, we analyzed the promoters of genes that were dsyregulated by Ad, scanning for over-represented transcription factor binding sites. AP-1, NF-κB, Oct1, Stat1, and Gut-enriched Krueppel-like factor (KLF4) are but a few of the more than 30 different transcription factors we identified as involved in the induction of the Ad innate immune response in MEFs (Hartman et al., 2007a). Supporting these observations and predictions, a recent array analysis by Granberg et al., (2006) found KLF4 to be upregulated within 2hpi of Ad infection of human IMR-90 cells. Future studies will be needed to verify these predictions in human cells, and/or in vivo.

Immune Cell Responses to Ad infection in vitro

In parallel with non-immune cells, multiple groups have demonstrated defense responses after Ad infection of more typical “immune cells”. For example, in cultured dendritic cells (DCs), various Ad vectors have demonstrated inflammatory effects, with studies mainly focused on the induction of NF-κB and upregulation of maturation markers (Hirschowitz et al., 2000; Korst et al., 2002; Varnavski et al., 2003). While Ad induced maturation of DCs was originally a matter of some dispute, the great majority of studies support the view that Ad infection activates DCs of both murine and human lineages (Dietz and Vuk-Pavlovic, 1998; Hirschowitz et al., 2000; Korst et al., 2002; Lyakh et al., 2002; Miller et al., 2003; Varnavski et al., 2003). Different serotypes of Ad appear to retain this capacity although there appears to be differences in the strength of induction from different serotypes. For example, chimpanzee serotype derived Ad vectors demonstrate significant upregulation of both IL-6 and Type I interferons (IFNs) in cultured human and mouse DCs, to a greater degree compared to their Ad5 counterparts (Varnavski et al., 2003). Significantly, this was the first display of a clear induction of Type I IFNα from Ad-infected myeloid DCs. In addition, induction of Type I IFNs by the chimpanzee AdC68 was shown to be independent of viral transcription and replication, and was also found to substantially inhibit vector derived transgene expression (Hensley et al., 2007; Hensley et al., 2005).

Likewise, multiple studies have found macrophages to be affected by adenoviral infection. Exposure of primary mouse alveolar macrophage cells to Ad virions resulted in the rapid induction of multiple cytokines including TNF-α, IL-6, MIP-2, and MIP-1α (Zsengeller et al., 2000). This effect was dependant on adenoviral uptake but not necessarily adenoviral nuclear transduction. Exposure of RAW264.7 mouse macrophages to Ad virions stimulated the rapid upregulation of TNF-α, an effect that was abolished when endosomal trafficking and acidification was prevented by pre-treating Ad infected cells with chloroquine (Zsengeller et al., 2000). A similar result was observed if endosomal escape of the virus was abolished by denaturing the virus at high temperatures prior to its exposure to cells (Trevejo et al., 2001; Zsengeller et al., 2000). Studies using an Ad mutant unable to lyse the endosome confirmed a lack of IRF3 induction following infection of primary mouse macrophages (Nociari et al., 2007). Together, these data indicate that elements of innate immune signaling against Ads in macrophages are dependent on endosomal trafficking and/or rupture.

In human peripheral blood mononuclear cells (PBMCs, innate cell precursors), Ad transduction also results in the transcriptional upregulation of the IL-6 and RANTES genes, as well as steady increases of GM-CSF, MIP-1α, and IL-8 transcripts over a 96-hr. window (Higginbotham et al., 2002). As with macrophages, these responses are maintained despite UV-inactivation of the virus. A recent study showing IL-6 and TNF-α induction in mouse macrophages after exposure to calcium chloride precipitated empty adenoviral capsids (Cerullo et al., 2007) was in contrast to work demonstrating that IRF3, nor other innate immune response genes, are induced in mouse macrophages after exposure to empty capsids (Nociari et al., 2007). Thus, in certain contexts, empty Ad capsids appear capable of inducing immune responses, but this may not be the case for all cell types. Future studies should focus on differentiating the importance of Ad capsids from the DNA content of the capsids in the elaboration of defensive immune responses, in different immune cell types.

Studies of Ad interactions with PRRs in vitro

The discovery and characterization of several different families of pathogen recognition receptors (PRRs) in the past decade have allowed investigators to probe the mechanism surrounding the initiation of cellular defense responses to adenoviral infection in vitro. To this end, we will present the relevant studies and findings involving adenoviral interactions with the Toll Like Receptor (TLR) and RLR (RIG-I Like Receptor) families of PRRs.

Role of TLR receptors in Ad induced innate immunity in vitro

A prominent group of PRRs that has recently been proven to play a pivotal role in the recognition of pathogens is the TLR family. Following TLR activation, a complex effector network is stimulated, resulting in a host defense response that is characterized by elaboration of a cascade of cytokines and chemokines, with subsequent influx of respondent professional immune cells that attempt to eliminate the invading pathogen. These receptors recognize many types of PAMPs found within viral and/or bacterial pathogens, including Triacyl lipopeptides (TLR1), zymosan (TLR2/6), viral dsRNA/poly(I:C) (TLR3), LPS (TLR4), bacterial flagellin (TLR5), viral ssRNA (TLR7/8), bacterial and viral CpG DNA (TLR9) and a profilin-like molecule (TLR11) (reviewed in (Lee and Kim, 2007b)). Mouse Mammary Tumor Virus (MMTV), Respiratory Synticial Virus (RSV), Human Papilloma Virus (HPV), Sendai Virus, and Measles Virus (Bieback et al., 2002; Compton et al., 2003; Czarneski et al., 2003; Haynes et al., 2001; Lopez et al., 2004; Lund et al., 2003; Means et al., 2003; Rassa et al., 2002; Yang et al., 2004) and now adenoviruses are but a few examples of viruses shown to precipitate an immune response through TLR dependent mechanisms.

Several lines of investigation have examined the in vitro involvement of the TLR family in Ad innate immunity. In one such study, Philpott et al., (2004) showed that Ad induced CD86 expression in dendritic cells (DCs) was independent of MyD88 signaling. In contrast, recent evidence has indicated that TLR9, an endosomally localized TLR shown to recognize CpG DNA, was required for the induction of various cytokines in response to Ad infection (see Figure 1). Peritoneal macrophages infected by HD-Ads also produced high levels of IL-6 in a TLR9 dependent manner (Cerullo et al., 2007). Furthermore, HD-Ad, co-precipitated with calcium phosphate (CaPO4) to achieve higher transduction of cells, induced MCP-1, IL-6 and TNF-α in peritoneal macrophages. However, only IL-6 and TNF-α inductions were TLR9 dependent, signifying a TLR9 independent mechanism for Ad induced MCP-1 production. In parallel, the induction of IL-6 and TNF-α following CaPO4 facilitated infections by ‘empty capsid’ Ads was also TLR9 independent, suggesting the requirement of Ad genomic DNA for induction of these cytokines through TLR9, as well as the presence of a capsid induced immune response independent of TLR9 signaling.

Figure 1. Intracellular and membrane bound sensors of Adenovirus.

Figure 1

Toll-Like Receptors (TLRs) are membrane bound Pattern Recognition Receptors (PRRs) that recognize conserved molecular structures known as pathogen-associated molecular patterns (PAMPs) found on invading pathogens. (A) Based on the cell type, Ad has been shown to elicit innate immune responses through TLR9 and MyD88 dependent pathways. Ad induced stimulation of these pathways results in rapid activation of over 30 transcription factors including NF-κB, and IRF3, as well as release of numerous cytokines and chemokines. (B) Ablation of either MyD88 or TLR9 does not fully suppress these innate responses, suggesting redundant systems, such as other TLR proteins, or adaptors such as TRIF, exist that stimulate immune response pathways that are independent of MyD88 and/or TLR9. (C) Intracellular sensors of dsDNA, such as RIG-I, are also likely PRRs involved in Ad induced innate signaling in various cell types resulting in similar activation of downstream pathways including activation of IRF3 and release of Type I IFNs. The effector molecules, downstream of the activated receptors, are currently unknown, but are likely shared with pathways known to be activated by other PAMPs. NFκB, nuclear factor kappa B; TRIF, TIR domain-containing adaptor protein inducing interferon beta; MyD88, myeloid differentiation primary response gene-88; IRF, interferon response factor; IFN, interferon; IL, interleukin; MCP, monocyte chemoattractant protein; RANTES, regulated on activation, normal T-cell expressed and secreted; MIP, macrophage inflammatory protein; IP, interferon gamma inducible protein; MAVS , mitochondrial antiviral signaling.

Iacobelli-Martinez and Nemerow, (2007) showed that Ad5 virus pseudotyped with the fiber from Ad16 (Ad5.F16, a CD46 receptor utilizing Ad) induced higher levels of Type I IFNs in peripheral blood mononuclear cells (PBMCs) than did the CAR utilizing Ad5. In a subsequent analysis, their data indicated that in HeLa cells, Ad5.F16 induced Type I IFN secretion and NF-κB activation to a greater level than Ad5, and that these responses were TLR9, but not TLR7 dependent. These results demonstrate that the route of Ad entry may substantially contribute to the type of innate immune response initiated by the TLR pathways. Finally, while Ad5 induction of Type I IFNs was mediated by TLR9 in a MyD88 dependent manner in plasmacytoid DCs (pDCs), it was induced in a MyD88 independent manner in conventional DCs (cDCs), and peritoneal macrophages (Basner-Tschakarjan et al., 2006; Zhu et al., 2007). These studies indicate a cell type specific role for TLR pathways in the innate immune response to Ad infection.

These cell-type specificities are mirrored by findings investigating the role of the TLR adaptor known as TRIF. TRIF is known to play a significant role in the downstream signaling of TLR3 and TLR4. Nociari et al., (2007) demonstrated that Ad induction of IL-6 is significantly diminished in TRIF deficient macrophages compared to wild-type counterparts, but that it is only partially diminished in TRIF deficient cDCs. Furthermore, the Ad-mediated induction of IFNα, IP-10, and RANTES in macrophages or cDCs is not affected by TRIF deficiency. We demonstrated the MyD88 and TRIF dependent Ad induction of cytokines in human A549 cells, bolstering the view that TLR adaptors figure prominently in the cellular immune responses to Ad vector infection (Hartman et al., 2007a; Hartman et al., 2007b).

Role of the RLR receptor RIG-I and Ad DNA in Ad induced innate immunity in vitro

The intracellular receptors RIG-I and MDA-5, serve as dsRNA detectors that together recognize the genomes of encephalomyocarditis virus, Newcastle disease virus, Sendai virus, influenza virus, and vesicular stomatitis virus independent of TLR regulated pathways (Gitlin et al., 2006; Kato et al., 2005; Kato et al., 2006). This recognition typically results in the recruitment of, or interaction with, the mitochondrial MAVS protein (also known as IPS1, VISA, or Cardif), stimulation of TANK-related kinases, and subsequent phosphorylation and activation of Type I IFN specific transcription factors known as interferon response factors, specifically IRF3 and IRF7 (Kumar et al., 2006). Recent evidence suggests that a similar TLR independent mechanism exists for the recognition of dsDNA, using Ad as the model system. As alluded to earlier, in murine bone marrow derived DCs (BMDCs) or bone marrow derived macrophages (BMDMs), Ad infection resulted in MyD88 and TRIF independent increases in CD86, CD40 and CD80 expressing cells, transcription of a variety of cytokines and chemokines, and activation of IRF3, presumably due to Ad capsid penetration of the endosome, and exposure of the Ad dsDNA genome to a cytoplasmic sensor (Nociari et al., 2007). A recent article by Cheng et al., (2007) describes, for the first time, a direct role for RIG-I in sensing intracellular dsDNA. Contrary to previous reports (Ishii et al., 2006; Kumar et al., 2006; Sun et al., 2006), the data in this study indicate that expression of dominant negative RIG-IC (or knocking-down MAVS using siRNA) completely suppresses IFNβ promoter activity following poly(dAT:dAT) or poly(I:C) treatment in the human hepatoma cell line Huh-7. In an experiment specific for Ad, Cheng et al., (2007) also showed that transfection of Ad DNA into Huh-7 cells containing a loss of function point mutation in RIG-I (Huh-7.5.1) also resulted in loss of IFNβ transcriptional induction. These results support evidence previously published by our lab implicating MAVS in Ad sensing. We specifically showed that shRNA mediated knock-down of MAVS protein expression in human lung epithelial A549 cells suppressed IFNβ promoter activity following Ad infection (Hartman et al., 2007a).

Thus, while there is growing evidence of the role played by TLR and RLR families in the sensing of adenoviral infection, future studies are needed to ascertain the identity of other TLR and RLR family members involved, the cell type and response specificities involved, as well as the identity of other PRR families that may participate in sensing adenoviruses.

AD INDUCED INNATE IMMUNITY IN VIVO

While in vivo cellular transduction is a potentially more complicated process than in vitro studies of cellular transduction, (i.e. involving putative binding or coating of Ad particles with several blood-borne factors prior to interactions with several different cell types) initial forays investigating the in vivo cellular response to Ads have nonetheless noted many similarities to in vitro investigations. Furthermore, the mechanism of innate immunity in vivo must also be considered in the context of different PRR repertoires present within different cell types in vivo. Thus, this portion of the review will focus on the roles of 1) extracellular factors, 2) cellular factors, and 3) membrane-bound PRRs (specifically TLRs) during Ad induced innate immune responses in vivo.

Role of extracellular molecules in Ad induced innate immunity in vivo

To recognize and defend against different pathogens in various settings throughout the body, mammals have evolved a wide array of extracellular PRRs that serve to identify molecular structures conserved among various pathogens, structures referred to as pathogen-associated molecular patterns (PAMPs). Multiple types of extracellular PRRs, including pentraxin and collectin family members, as well as protein members of the complement pathways, both classical (C1q, C4) and alternative (Factor B, Factor D), opsonize pathogens to facilitate their phagocytosis and destruction, many times via interactions with cellular receptors (C1qR, CR1, CR2, CR3, CR4). In the context of adenoviral infection in vivo, Ad has been determined to interact with surfactant and complement in addition to interacting with pre-existing immunoglobulins.

Since breaching the keratinized epithelium found within the skin is a challenge for most viruses, entry is usually achieved by an alternate route, most notably through mucosal membranes. Since Group C human Ads primarily infect the upper respiratory tract, it was parsimonious to postulate an Ad interaction with surfactant, a family of C-type collections (C-type lections) primarily produced by the epithelial cells of the lung. In a 1999 study, Harrod et al., (1999) demonstrated greater inflammation and decreased lung clearance after intra-tracheal administration of FG-Ad5 in Surfactant-A (SP-A) deficient mice, as compared to their wild-type counterparts. Ad infected, surfactant deficient mice displayed greater levels of epithelial damage, higher numbers of infiltrating inflammatory cells, and higher levels of IL-6, TNF-α, and IL-1β in bronchoalveolar lavage fluid. Interestingly, increased levels of Ad vector derived DNA were found in the lungs of Ad infected SP-A deficient mice, relative to Ad infected normal mice. Other studies have noted this interaction, and documented the disruption of surfactant homeostasis after administration of Ad in a murine lung model (Ross et al., 1995; Zsengeller et al., 1997). Thus, SP-A appears to block Ad vectors in vivo, reducing their inflammatory profile through unproven but likely opsonizing effects.

One of the benefits of Ad based gene transfer vectors is their ability to successfully transduce large amounts of hepatic tissue following intravenous injection (Einfeld et al., 2001; Nicklin et al., 2005; Nicol et al., 2004; Shayakhmetov et al., 2005a; Shayakhmetov et al., 2004). This requires levels of virus several logs higher than would normally be found during a wild-type infection, a situation that likely amplifies Ad interactions with humoral proteins and PRR’s that have evolved to respond to much lower viral concentrations. This model, however, also allows detection of host-pathogen interactions that might not be detectable (but no less relevant) during wild-type infections. For example, our group and others have shown that Ad4 or Ad5 can activate the complement system in the presence or absence of pre-existing anti-Ad immunoglobulins through either the classical pathway, or through direct interactions with proteins of the alternative pathway respectively (Cichon et al., 2001; Jiang et al., 2004). Shayakhmetov et al., (2005a) demonstrated that fiber knob interactions with both blood factor IX and C4 binding protein were possible in vitro, and that ablating this interaction reduced serum levels of IL-6 and IFNγ 6hpi of this mutant virus in vivo.

In array based analyses of Ad induced liver transcriptome responses (comparing Ad injected wild-type and complement factor C3 (C3-KO) mice) Kiang et al., (2006) reported that of all liver genes dsyregulated 6 hours after Ad injection, 3 in 5 Ad upregulated genes, and 1 in 5 Ad downregulated genes were in part complement dependent for these transcriptional responses. Like those identified in MyD88 deficient mice (see below), the Ad induced, complement dependent genes identified in this analysis were largely representative of NF-κB, apoptosis, Th1, and IL-6 signaling pathways. This may indicate an interaction between complement and TLR pathways that has yet to be identified, but of potential relevance to several pathogens in addition to Ads per se. Supporting this notion, a subset of Ad induced plasma cytokines were also dependent on MyD88, and complement functionality. These include: KC, IL-5, G-CSF, GM-CSF, and IL-6. Future studies dissecting potential Ad complement TLR system interactions, as well as efforts to capitalize on this knowledge to improve the safety profile of Ad vectors in vivo are now underway.

The host has several other humoral innate immune mechanisms poised to rapidly identify, sequester, and neutralize low levels of pathogens not previously encountered. These “hair-triggered” guardians include non-specific, pre-existing “natural” immunoglobulins, lysozymes, C-reactive protein, collectins, pentraxins, serum amyloid, as well as mannose binding proteins. These molecules can bind pathogens and prevent their entry into host cells as well as opsonize them to facilitate their phagocytic digestion by innate immune cells. However, the roles of these systems in Ad mediated immune responses remain currently unknown.

Finally, most viruses, including Ads, elicit neutralizing, Ad specific antibodies after primary inoculation into a host (Jooss and Chirmule, 2003; Malkevitch et al., 2003; Santra et al., 2005). Binding of Ad vectors by specific immunoglobulins facilitates rapid clearance of the vector by cellular elements of the innate immune response, likely including Kupffer cells, and other cells of the reticuloendothelial system (RES), as well as activation of the classical arm of the complement pathway (Cichon et al., 2001). Recently, it has been shown that that Ad vectors may be more toxic in the presence of anti-Ad antibodies than in their absence, the etiology of which is unclear at the moment (Varnavski et al., 2005; Varnavski et al., 2002).

The limited studies preformed thus far indicate that adenoviral virions interact with and are opsonized by surfactant-A in the lungs and by complement in the bloodstream. Opsonization by surfactant facilitates viral clearances and results in a less inflammatory phenotype while opsonization by complement results in a more inflammatory phenotype in an intravenous setting. In this same vain, binding by pre-existing immunoglobulins results in a more rapid uptake by the RES and greater inflammatory effects in vivo. Thus, the effect of opsonization by extracellular factors depends on the type of extracellular factor involved and the context of infection. While only a few extracellular systems have been studied, it is likely that adenoviral particles interact with multiple types of extracellular PRRs to generate the particular innate and inflammatory responses observed in vivo.

Cellular elements of the in vivo innate immune response to Ads

The first investigations of the in vivo innate immune response to adenoviral vectors proceeded from gene therapy based approaches that used intravenous administration of type 5 adenoviral vectors to highly infect mouse hepatic tissue. Several studies have shown that Ad transduction into the murine liver results in the upregulation of many of the same interferon-related genes found in vitro (e.g. MCP-1, IP-10, RANTES, MIP-2, etc.) reaching peaks at 6–12 hours post-injection with a return to near-baseline by 24 hours post injection. Infection with UV inactivated and HD-Ad vectors showed similar gene induction patterns, however, induction of a second phase response at 5dpi and 7dpi required Ad derived viral gene transcription (Liu et al., 2003; Zaiss et al., 2002). Our studies using a microarray based approach to analyze in vivo Ad induced immune responses demonstrated that a large portion of liver tissue derived transcripts (∼15%) are affected by adenoviral infection, reaching peak levels at 6hpi. While many of the Ad affected genes belong to innate immune families, other groups of Ad-dysregulated genes include RNA regulatory, apoptosis, lysosome, and endocytic families of genes (Hartman et al., 2007b).

Whether or not Ad induced immune activation is correlated with hepatocyte transduction is currently unresolved, and use of hepatically detargeted Ad vectors has not provided much clarification. While some groups have reported a lack of hepatic transduction with penton RGD and fiber CAR binding double mutants in certain strains of mice and rats, others have reported high levels of in vivo hepatic transduction from these same viruses in other strains of mice, rats and in primates (Einfeld et al., 2001; Martin et al., 2003; Ni et al., 2005; Nicol et al., 2004; Smith et al., 2003a; Smith et al., 2003b). Recent evidence suggests that the HSPG binding motif present in the Ad5 fiber is sufficient for in vivo hepatic transduction in the absence of both CAR and RGD motifs (Kritz et al., 2007).

The emerging consensus favors the importance of the HSPG motif and potentially the size of the fiber used relative to Ad upregulated innate immune response genes. For instance, in DBA/2 mice infected with RGD-ablated vectors, significantly lower transcription levels of IP-10 and MIP-2 were noted, as compared to those receiving a similar dose of wild-type Ad vectors (Liu et al., 2003). In C57BL/6 mice, injection with short-shafted Ad vectors, which display fibers without HSPG motifs, also led to decreased liver transduction and significantly lower transcription levels of MIP-2 and IL-1α, in comparison to Ad vectors displaying wild type fibers (Shayakhmetov et al., 2004). Intriguingly, in mice injected with these vectors, IP-10 upregulation was similar but delayed, leading the authors to speculate that induction of this gene is not dependent upon hepatocyte transduction per se. Elimination of resident macrophages by pretreatment of mice with agents such as gadolinium chloride does not change upregulation of these genes in vivo, but in fact accentuates their upregulation, potentially through the increased hepatic transduction afforded by macrophage depletion (Liu et al., 2003). These findings provide evidence for the importance of stromal cell transduction in Ad induced innate immune responses.

The systemic elaboration of several cytokines and chemokines after Ad injection accompanies the induction of innate immune genes. The cellular origin of these systemic cytokines is not fully known, although evidence obtained thus far suggests prominent roles played by innate immune cells specifically, such as macrophages and dendritic cells. The importance of macrophages is most easily demonstrated in studying Ad’s natural tropism of the respiratory tract. In an in vivo study, Zsengeller et al., (2000) showed rapid accumulation of Ad5 vectors in alveolar macrophages 10 minutes after vector administration in the murine upper respiratory tract. This effect was strongly associated with the induction of inflammatory cytokines and chemokines (such as IL-6, TNF-α, MIP-2, and MIP-1α) produced by alveolar macrophages.

Other recent studies have shown the importance of dendritic cells in the Ad mediated induction of cytokines in vivo. Using intravenous injection of fiber modified Ad vectors, Koizumi et al., (2007) found that systemic induction of IL-6 (and to a lesser extent IL-12) correlated with high vector uptake by (but low transduction of) splenic cells. When several splenic cell types were assessed after Ad injection, only conventional B220-CD11c+ DCs were found to have significant induction of IL-6 and IL-12 transcription. However, macrophage populations were not assessed in this study, thus it is unclear if resident or trafficking nonresident macrophage populations are also responsible for cytokine production in the spleen.

Kupffer cells are the primary macrophages resident in the liver, and have the primary function of sequestering pathogens that enter the liver vasculature. Kupffer cells are known to play a large role in the sequestration of intravenously injected Ad vectors, thereby preventing hepatocyte transduction unless saturated by overwhelming Ad doses. Depletion of Kupffer cells prior to Ad injection results in correction of the non-linear dose threshold effect, i.e. lower dosage Ad infusions are more capable of efficiently transducing greater numbers of hepatocytes (Lieber et al., 1997; Worgall et al., 1997; Ziegler et al., 2002). This improved ability to transduce hepatocytes is accompanied by a decrease in Ad induced endothelial cell activation and/or cytokine release (Schiedner et al., 2003; Snoeys et al., 2006). Kupffer cell sequestration of Ad particles has also been confirmed in baboons, with injection of 1.2 × 1012 particles/kg providing evidence of both Kupffer cell sequestration and cytokine induction (IL-6) but also some limited hepatocyte transduction. A 10 fold higher dose (1.2 × 1013 particles/kg) resulted in widespread liver transduction by the vector, but this dose was also lethal to the treated animals.

In addition to Kupffer cells, endothelial cells also sequester large amounts of Ad vectors following injection of higher doses in primates. These patterns of Ad vector sequestration and toxicity mimic those noted in mice, at near equivalent dosages per kilogram (Morral et al., 2002; Tao et al., 2001). Though technically not cells, blood platelets may figure prominently in these responses. It has been shown that Ads can infect platelets in vitro, and mice intravenously injected with Ad vectors can be found to contain Ads within their platelets as well (Faraday et al., 1999; Gillitzer et al., 2005; Stone et al., 2007). The significance of Ad interactions with platelets (i.e.: upon subsequent Ad induced innate or adaptive immune responses) awaits future investigations.

Finally, several groups have shown that coating of Ad using Poly-ethylene glycol (PEG) prior to injection can effectively reduce the production of several inflammatory cytokines such as IL-6, IL-12, and TNF-α, as well as Ad induced thrombocytopenia. Studies have also shown that “PEGylation” results in a reduced uptake of Ad vectors by Kupffer cells and macrophages in vitro (Croyle et al., 2002; Croyle et al., 2005; O'Riordan et al., 1999) These complexities point to future studies required to dissect out and assess the significance of the multi-faceted interactions occurring between Ads and blood proteins, platelets, macrophages, endothelial cells and respective parenchymal cells.

Role of cellular TLRs in Ad induced innate immunity in vivo

Unlike in vitro studies, to our knowledge there have been no published studies to date investigating the effect of other cellular PRRs (such as RLR or NOD family members) in Ad immunity in vivo. This area constitutes a gap in adenoviral immunity which future studies should address. As with in vitro investigation of TLRs, there have been few studies directly assessing the interaction of Ad and TLRs.

The first study investigating Ad interaction with TLRs in vivo was actually carried out before the discovery of TLRs. In 1991, Ginsberg et al., (1991) investigated the molecular pathogenesis of adenoviral induced pneumonia in mice, using several inbred mouse strains. They discovered that although adenovirus did not replicate in mouse lungs, pathogenic pneumonia would develop due to the inflammation induced by the high concentration of viral inoculum used. Curiously, they noted the highest level of inflammation in C57BL/6N and CBA/N mice and the lowest levels of inflammation in C57BL/10ScN mice, a strain with an endogenous deletion of TLR4 (Poltorak et al., 1998). Eight years later, Thorne et al., (1999) found that Ad infected C3HeB/FeJ mice (wild-type TLR4) elicited much greater Ad-induced lung inflammatory responses when compared to congenic, C3H/HeJ (TLR4 defective) mice. These studies implicate TLR4 as being involved in the innate immune response to Ad in the lungs, although the molecular characterizations of these interactions remain unexplored. In contrast, a different study showed that C3H/HeJ (TLR4 defective) mice intravenously injected with high doses of Ad exhibited some inflammatory reactions noted after high dose intravenous Ad injections of wild-type mice, although these limited studies did not include tandem comparisons to TLR4 wild-type congenic mice (Lieber et al., 1997).

Using a broader approach, our group recently completed a set of experiments aimed at delineating the role of the critical TLR adaptor MyD88 subsequent to intravenous administration of Ad vectors. For example, we defined the MyD88 dependent aspects of the Ad induced liver, transcriptome profile in vivo (Hartman et al., 2007b). We found that 15% of all assessable gene probes represented on the array were significantly dysregulated by Ad in the livers of wild type animals. In comparison, livers of Ad treated, MyD88 deficient mice exhibited dysregulation of only 4.4% of assessable gene probes. In this analysis, a significant number of TLR, MAPK, apoptosis and complement activation genes were identified as MyD88 dependent, Ad upregulated genes. The Ad induction of plasma cytokines and chemokines, such as KC, G-CSF, IL-6 and MCP-1 were also found to be partially MyD88 dependent at 1hpi, and Ad inductions of plasma MCP-1, MIP-1α, IL-5, IL-6, IL-12(p40), G-CSF, GM-CSF and RANTES were partially MyD88 dependent at 6hpi. Indirectly supporting this work, Shayakhmetov et al., (2005b) showed reduced liver toxicity as measured by ALT, as well as reduced plasma levels of IL-1β and IL-6 following injection of Ad into IL-1R1 deficient mice, a result also likely mediated by MyD88, as the IL-1R also utilizes MyD88 as an intra-cellular signaling adaptor.

Validating the importance of innate immune responses in directing downstream adaptive immune responses to pathogens, we have also demonstrated that in Ad injected MyD88 KO mice, anti-Ad IgG and transgene-specific IFNγ T-cell responses were significantly lower in Ad injected MyD88 deficient mice, as compared to Ad injected controls. TLR9 has also been implicated as a specific TLR involved in the immune response to Ad in vivo. In TLR9 deficient mice, induction of IL-6, IL-12 at 6hpi and IFNα at 12hpi, has been shown to be partially dependent on TLR9. Notably, the induction of KC and MCP-1 was not dependent on TLR9, and verified these cytokines’ dependence upon MyD88 functionality (Cerullo et al., 2007). In vitro work by a different group demonstrated that Ad infection induced IFN-α in a TLR9-MyD88 dependant fashion in pDCs but not other cell types (Zhu et al., 2007).

At present, the collective knowledge surrounding Ad interactions with alternate TLRs or intracellular receptors in vivo is relatively lacking. As mentioned, it is likely that other TLRs, along with TLR9, are also required for the complete induction of the innate immune response to Ads, based on the induction of multiple cytokines in TLR9 deficient mice, and the continued dysregulation of Ad induced immune pathways in Ad injected, MyD88 deficient animals. Evidence in vitro indicates that the RIG-I/MAVS pathway could be involved in these responses. These findings do support the notion that exploring and understanding the nuances of Ad vector interactions with the innate immune system can have far reaching implications, in this instance providing justification for studies attempting to understand how TLR responses to Ad vectors impact upon Ad vector efficacy in vaccine applications. Finally, Ad interactions with other cellular PRRs have not yet been investigated in vivo, further underscoring the large gaps in our knowledge of the mechanisms underlying adenoviral innate immunity in vivo.

Conclusion

Modified Ads are commonly used for a wide-range of scientific and therapeutic endeavors, such as vaccine platforms, oncolytic agents, or gene transfer vectors for targeted applications both in vitro and in vivo. This wide applicability makes understanding the innate immune response to these viruses particularly important for improving efficacy in each of these types of endeavors. This review highlights that the nature of Ad induced immune responses are complex, despite the limitations of current technologies used to define these responses. As improved technologies emerge, and more sophisticated animal models become available (i.e.: those with multiple gene modifications simultaneously present in a single animal) these complexities will likely broaden.

Collectively, the data reviewed in this manuscript begins to identify the scope of the Ad induced innate immune response, as it involves extracellular, intracellular, and membrane bound receptors, that make up a complex network of immune response pathways that have co-evolved to rapidly respond to invading pathogens such as Ads. These responses vary based on the type of entry Ad initiates, the type of DNA incorporated into the virion, as well as the extracellular milieu encountered by the virion before infection.of various cell types. As a consequence, these variables are both capable of limiting the efficacy of Ad based gene transfer vectors, as well as augmenting the efficacy of Ad based vaccine vectors. Furthermore, as pseudotyped Ads, as well as alternative serotype Ad based vectors become more widely utilized, it will also be likely that each of the vectors will not only offer unique insights, but also encounter new limitations , based in large part on their altered interactions with the innate immune system.

Because of these considerations, the Ad vector community specifically, and the gene transfer community at large, must continue to vigorously identify, investigate, and understand the innate immune systems that may be triggered upon introduction of foreign transgenes by any gene transfer platform. Without these types of investigations, future gene therapy trials may again encounter untoward or unforeseen outcomes attributable to innate immune responses, while therapeutic vaccination and cancer therapy trials utilizing vectors may show decreased efficacy secondary to misdirected and/or vector-limited immune responses. Continued efforts to understand the innate immune response will lead to greater wisdom in the use of all gene transfer based vectors, enabling future investigators to more effectively and pro-actively manipulate the innate immune system, to either minimize the toxicity, or accentuate the immune response desired after Ad vector mediated gene transfer specifically, and/or via other vectors generally.

Table 1.

Summary of in vitro analyses of Ad induced responses in infinite and finite lifespan cells derived from various tissues from both human and murine species

Cell Line Cell Type Species Type
HeLa Adenocarcinma epithelial cells Human Immortal
M21 Melanoma derived Human Immortal
HB2 Mammary epithelial cells Human Immortal
A549 Lung carcinoma cell line Human Immortal
Huh-7 Hepatoma cell line Human Immortal
PBMCs Peripheral blood mononuclear cells Human Primary Culture
Myeloid DCs Dendritic cells derived from PBMCs Human Primary Culture
IMR-90 Lung fibroblast Human Primary Culture
RAW264.7 Ascites derived macrophage Murine Immortal
REC Renal epithelial cells Murine Immortal
TGP61 Pancreatic insulinoma Murine Immortal
BMDCs Bone marrow derived dendritic cells Murine Primary Culture
MEF Mouse embryonic fibroblasts Murine Primary Culture
PM Peritoneal macrophages Murine Primary Culture
pDCs Plasmacytoid DCs Murine Primary Culture
cDCs Conventional DCs Murine Primary Culture
KC Hepatic Kupffer cells Murine Primary Culture
Cell Line Time Frame Ad Induced Immune Response References
HeLa 30mpi; 2, 16, 48hpi ERK and MKK activation, IL-8 transcription (30mpi); NFκB activation (6hpi); RANTES (16hpi), IFNβ induction (48hpi) Bruder et al., 1997; Bowen et al., 2002; Iacobelli-Martinez et al., 2007
*HeLa 12, 32, 48hpi Array analysis revealed max induction of IFN regulated genes at 12hpi Dorn et al., 2005
M21 30hpi Array analysis finding 89 gene changes common to Ad5, Ad5-RGD mutant and Ad5/3 chimera infection Volk et al., 2005
HB2 24hpi, 72hpi Array analysis finding induction of 24 (24hpi), and 96 (72hpi) genes including IL-8, IL-6, NFκB Scibetta et al., 2005
A549 20mpi, 16hpi, 24hpi Activation of MAPK(20mpi), NFκB, AP-1 and Oct-1 (16hpi); MAVS, MyD88, TRIF required for IFNβ promoter activity Hartman et al., 2007; Tamanini et al., 2006
Huh-7 4, 12, 24hpi Induction of IFNβ transcript; array approach: WT-Ad5, HD-Ad5, FG-Ad5 shows differentially modulated pathways Cheng et al., 2007, Martina et al., 2007
PBMCs 0 – 96hpi TNFα, IL-1β, GMCSF - gradual increase; IL-6, RANTES- Max at 30hpi; IL-8- Max at 72hpi; 24hpi IFNα induction (24hpi) Higginbotham et al., 2002, Iacobelli-Martinez et al., 2007
Myeloid DCs 24, 48hpi IFNα induction; Increased DC surface markers CD40, CD80, CD86 Varnavski et al., 2003
IMR-90 2hpi, 24hpi Array analyses identifying KLF4 induction (2hpi) and 3-fold induction of 4.2% of assessed genes (24hpi) Stilwell et al., 2004; Granberg et al., 2006
RAW264.7 2hpi TNFα release, blocked by chloroquine and heat denaturation Zsengeller et al., 2000
REC 10mpi; 6hpi PI3K dependent Akt activation; ERK and p38 activation (10mpi); IP-10 induction, NFκB activation (6hpi) Tibbles et al., 2002; Borgland et al., 2000; Liu et al., 2003/2005
TGP61 1hpi, 6hpi Induction of MIP-2 (1hpi), RANTES, IP-10 (6hpi) transcription Muruve et al., 1999
BMDCs 18hpi, 36hpi, 48hpi IFNα, IFNβ, IL-6, IL-12(p40), IL-1β induction; Increased DC surface markers CD86, CD54,CD40; IRF3 activation Hirschowitz et al., 2000; Korst et al., 2002; Lyakh et al., 2002;
5hpi Induction of TNFα, IFNβ, CCL5, CCL1, CCL4, and IP-10 transcription Cerullo et al., 2007, Nociari et al., 2007
MEF 16hpi Array analysis revealed induction of 22% of assessed genes, induction of 30 TFs and MIP1a, IL-6, KC and RANTES Hartman et al., 2007
PM 6hpi, 18hpi IL-6, TNFα induction (6hpi); IFNα, IFNβ induction (18hpi) Cerullo et al., 2007; Zhu et al., 2007
pDCs 18hpi IFNα, IFNβ induction Zhu et al., 2007
cDCs 18hpi IFNα, IFNβ induction Zhu et al., 2007
KC 18hpi IFNα, IFNβ induction Zhu et al., 2007
*

Wild-type Ad12 was used to infect HeLa cells as opposed to wild-type Ad5 or FG-Ad5

Table 2.

Summary of in vivo analyses of Ad induced innate immune responses

Strain Genetic Background Challenge Virus Dose Time Frame Ad induced immune resposne Reference
BALB/c Wild-Type Intratracheal FG-Ad5LacZ 4 × 1010 OPU (v.p.) 0.5, 3, 6 hours Induction of IL-6, TNFα, MIP2, MIP1α in BALF; inductio of trascripts in AM Zsengeller et al., 2000
C57BL/6 Wild-Type Intravenous FG-Ad5[E1-] 5 × 107 p.f.u. 3 days cDNA array approach identified several induced transcripts Haberberger et al., 2000
C57BL/6 Wild-Type Intravenous FG-Ad5/9 short shaft 1 × 1011 v.p. 0.5, 6, 24hpi Reduced MIP2, IL1α transcripts; delayed IP-10 induction; Reduced plasma TNFα, IL-6, MCP1, IFNγ levels Shayakhmetov et. al, 2004
C57BL/6 Wild-Type Intravenous FG-Ad5-Fiber Modified (7K) 1 × 1011 v.p. 3, 24, 48hpi Induction of IL-6 correlates with splenic transduction Koizumi et al., 2007
C57BL/6 Wild-Type Intravenous HD-Ad; FG-Ad5βGal - PEGylated 8 × 1010 v.p. 6hpi Reduced serum IL-6, IL-12, TNFα induction; Reduced thrombocytopenia Croyle et al., 2005
C57BL/6 Wild-Type Intravenous FG-Ad5LacZ 1.5 × 1011 v.p. 1, 6, 24hpi Induction of serum KC (1hpi), IL-5, GCSF, GM-CSF, IL-6, IFNy, IL-1β, IL-12(p40) RANTES; Array approach identifies
15–19% of assessable liver genes dysregulated at 6hpi
Kiang et al., 2006; Hartman et al., 2007
C57BL/6 Wild-Type Intravenous FG-Ad5 + TNFα antibody 1 × 1011 v.p. 0.5, 6, 24hpi Reduced plasma IL-6 levels (6hpi) Shayakhmetov et. al, 2005
C57BL/6 Wild-Type Intravenous FG-Ad5 + IL-1α/β antibodies 1 × 1011 v.p. 0.5, 6, 24hpi Reduced IL-1α, IL-1β transcripts; reduced plasma IL-6 (6hpi); reduced ALT Shayakhmetov et. al, 2005
C57BL/6 Wild-Type Intravenous FG-AdLacZ 2 × 109 p.f.u. 6hpi, 12hpi Induction of serum IFNα, IL-6 Zhu et al., 2007
C57BL6 Wild-Type Intravenous FG-Ad5mut no C4/FIX binding 1 × 1011 v.p. 6hpi Reduced serum IL-6 and IFNγ induction Shayakhmetov et. al, 2005
DBA/2 Wild-Type Intravenous FG-Ad5Luc 1 × 1011 v.p. 1hpi - 7dpi Induction of biphasic MIP1β, MIP1α, MIP-2, MCP-1, IP-10, TNFα trancripts, induction of sustained RANTES
transcript; Neutrophil Infiltration (1hpi)
Liu et al., 2003
DBA/2 Wild-Type Intravenous FG-Ad5LacZ 109 – 1011 v.p. 1, 6, 24hpi Dose dependent induction of RANTES, MIP1β, MIP2, MCP-1, IP-10 transcripts Zaiss et al., 2002
DBA/2 Wild-Type Intravenous FG-Ad5[E1-] - RGD mutant 1 × 1011 v.p. 1, 6, 24hpi Reduced IP-10 and MIP-2 transcript induction; reduced Neutrophil Infiltration Liu et al., 2003
Several Wild-Type Intranasal Wild Type Ad5 1 × 1010 p.f.u. 1 – 7 days Induction of lung and plasma TNFα, IL-1 and IL-6; induction of IL-1β transcript in lung Ginsberg et al., 1991
C57BL6/129s TLR9 Deficient Intravenous HD-Ad5LacZ 5 × 1012; 1 × 1013 v.p./kg 6hpi Plasma IL-6, IL-12 reduced Cerullo et al., 2007
C57BL6/129s TLR9 Deficient Intravenous FG-Ad5LacZ 2 × 109 p.f.u. 12hpi Reduced IFNα Zhu et al., 2007
C3H/HeJ TLR4 Deficient Intratracheal FG-Ad5LacZ 1 × 1010 v.p. 3, 12, 24hpi Reduced neutrophil infiltration, TNFα, IL-6 in BALF compared to control C3HeB/FeJ Thorne et al., 1999
C3H/HeJ TLR4 Deficient Intravenous FG-Ad5hAAT 1 × 1010 TU 0 – 96hpi Induction of TNFα and IL-6, however not compared to control C3HeB/FeJ Lieber et al., 1997
C57BL/10ScN TLR4 Deficient Intranasal Wild Type Ad5 1 × 1010 p.f.u. 1 – 7 days Lower overall inflammatory response Ginsberg et al., 1991
129/J SP-A Deficient Intratracheal FG-Ad5Luc 1 × 109 p.f.u. 6, 24hpi Increased TNFα, IL-6, MCP1, MIP1α, MIP2 in lung homogonate; increased TNFα, IL-1β, IL-6 in BALF Harrod et al., 299
C57BL6/129s MyD88 Deficient Intravenous FG-Ad5LacZ 1.5 × 1011 v.p. 1hpi, 6hpi Array analysis revealed almost 50% Ad dysregulated genes are MyD88 dependent; plasma KC (1hpi), GM-CSF,
GCSF, IL-5, IL-6, IL-12(p40), MIP1α, MCP1, RANTES (6hpi) reduced
Hartman et al., 2007
C57BL6/129s MyD88 Deficient Intravenous FG-Ad5LacZ 2 × 109 p.f.u. 12hpi Reduced IFNα Zhu et al., 2007
C57BL6/129s IL-1R Deficient Intravenous FG-Ad5βGal 1 × 1011 v.p. 0.5, 6 24hpi Altered IL-1α, TNFα, MIP2 gene expression, Reduced plasma IL-1β, IL-6, ALT levels Shayakhmetov et. al, 2005
129Sv IFNαβR Deficient Intravenous FG-Ad5LacZ 2 × 109 p.f.u. 6hpi Prolonged transgene expression, reduced neutralizing Ab titer Zhu et al., 2007
C57BL6/129s C3-Deficient Intravenous FG-Ad5LacZ 1.5 × 1011 v.p. 6hpi Array analysis revealed 60% Ad upregulated, and 20% Ad downregulated genes are complement dependent;
reduced plasma KC (1hpi), IL-5, GCSF, GM-CSF and IL-6 compared to wild-type controls
Kiang et al., 2006

FG, First Generation; HD, Helper-Dependent; p.f.u., plaque forming units; v.p., viral particles; TU, transducing units; OPU, optical particle units

Acknowledgements

AA was supported by NIH grants DK069884 and CA078673, and the Osteopathic Heritage Foundation.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  1. Applequist SE, Wallin RP, Ljunggren HG. Variable expression of Toll-like receptor in murine innate and adaptive immune cell lines. Int Immunol. 2002;14(9):1065–1074. doi: 10.1093/intimm/dxf069. [DOI] [PubMed] [Google Scholar]
  2. Basner-Tschakarjan E, Gaffal E, O'Keeffe M, Tormo D, Limmer A, Wagner H, Hochrein H, Tuting T. Adenovirus efficiently transduces plasmacytoid dendritic cells resulting in TLR9-dependent maturation and IFN-alpha production. J Gene Med. 2006;8(11):1300–1306. doi: 10.1002/jgm.964. [DOI] [PubMed] [Google Scholar]
  3. Beutler B, Rietschel ET. Innate immune sensing and its roots: the story of endotoxin. Nat Rev Immunol. 2003;3(2):169–76. doi: 10.1038/nri1004. [DOI] [PubMed] [Google Scholar]
  4. Bieback K, Lien E, Klagge IM, Avota E, Schneider-Schaulies J, Duprex WP, Wagner H, Kirschning CJ, Ter Meulen V, Schneider-Schaulies S. Hemagglutinin protein of wild-type measles virus activates toll-like receptor 2 signaling. J Virol. 2002;76(17):8729–36. doi: 10.1128/JVI.76.17.8729-8736.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Borgland SL, Bowen GP, Wong NC, Libermann TA, Muruve DA. Adenovirus vector-induced expression of the C-X-C chemokine IP-10 is mediated through capsid-dependent activation of NF-kappaB. J Virol. 2000;74(9):3941–3947. doi: 10.1128/jvi.74.9.3941-3947.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Bowen GP, Borgland SL, Lam M, Libermann TA, Wong NC, Muruve DA. Adenovirus vector-induced inflammation: capsid-dependent induction of the C-C chemokine RANTES requires NF-kappa B. Hum Gene Ther. 2002;13(3):367–79. doi: 10.1089/10430340252792503. [DOI] [PubMed] [Google Scholar]
  7. Bruder JT, Kovesdi I. Adenovirus infection stimulates the Raf/MAPK signaling pathway and induces interleukin-8 expression. J Virol. 1997;71(1):398–404. doi: 10.1128/jvi.71.1.398-404.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Cerullo V, Seiler MP, Mane V, Brunetti-Pierri N, Clarke C, Bertin TK, Rodgers JR, Lee B. Toll-like receptor 9 triggers an innate immune response to helper-dependent adenoviral vectors. Mol Ther. 2007;15(2):378–385. doi: 10.1038/sj.mt.6300031. [DOI] [PubMed] [Google Scholar]
  9. Cheng G, Zhong J, Chung J, Chisari FV. Double-stranded DNA and double-stranded RNA induce a common antiviral signaling pathway in human cells. Proc Natl Acad Sci U S A. 2007;104(21):9035–9040. doi: 10.1073/pnas.0703285104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Cichon G, Boeckh-Herwig S, Schmidt HH, Wehnes E, Muller T, Pring-Akerblom P, Burger R. Complement activation by recombinant adenoviruses. Gene Ther. 2001;8(23):1794–1800. doi: 10.1038/sj.gt.3301611. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Compton T, Kurt-Jones EA, Boehme KW, Belko J, Latz E, Golenbock DT, Finberg RW. Human cytomegalovirus activates inflammatory cytokine responses via CD14 and Toll-like receptor 2. J Virol. 2003;77(8):4588–4596. doi: 10.1128/JVI.77.8.4588-4596.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Croyle MA, Chirmule N, Zhang Y, Wilson JM. PEGylation of E1-deleted adenovirus vectors allows significant gene expression on readministration to liver. Hum Gene Ther. 2002;13(15):1887–900. doi: 10.1089/104303402760372972. [DOI] [PubMed] [Google Scholar]
  13. Croyle MA, Le HT, Linse KD, Cerullo V, Toietta G, Beaudet A, Pastore L. PEGylated helper-dependent adenoviral vectors: highly efficient vectors with an enhanced safety profile. Gene Ther. 2005;12(7):579–87. doi: 10.1038/sj.gt.3302441. [DOI] [PubMed] [Google Scholar]
  14. Czarneski J, Rassa JC, Ross SR. Mouse mammary tumor virus and the immune system. Immunol Res. 2003;27(2–3):469–80. doi: 10.1385/IR:27:2-3:469. [DOI] [PubMed] [Google Scholar]
  15. Dietz AB, Vuk-Pavlovic S. High efficiency adenovirus-mediated gene transfer to human dendritic cells. Blood. 1998;91(2):392–8. [PubMed] [Google Scholar]
  16. Dorn A, Zhao H, Granberg F, Hosel M, Webb D, Svensson C, Pettersson U, Doerfler W. Identification of specific cellular genes up-regulated late in adenovirus type 12 infection. J Virol. 2005;79(4):2404–12. doi: 10.1128/JVI.79.4.2404-2412.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Einfeld DA, Schroeder R, Roelvink PW, Lizonova A, King CR, Kovesdi I, Wickham TJ. Reducing the native tropism of adenovirus vectors requires removal of both CAR and integrin interactions. J Virol. 2001;75(23):11284–11291. doi: 10.1128/JVI.75.23.11284-11291.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Faraday N, Rade JJ, Johns DC, Khetawat G, Noga SJ, DiPersio JF, Jin Y, Nichol JL, Haug JS, Bray PF. Ex vivo cultured megakaryocytes express functional glycoprotein IIb-IIIa receptors and are capable of adenovirus-mediated transgene expression. Blood. 1999;94(12):4084–92. [PubMed] [Google Scholar]
  19. Gillitzer A, Peluso M, Laugwitz KL, Munch G, Massberg S, Konrad I, Gawaz M, Ungerer M. Retroviral infection and selection of culture-derived platelets allows study of the effect of transgenes on platelet physiology ex vivo and on thrombus formation in vivo . Arterioscler Thromb Vasc Biol. 2005;25(8):1750–1755. doi: 10.1161/01.ATV.0000172659.01157.c6. [DOI] [PubMed] [Google Scholar]
  20. Ginsberg HS, Moldawer LL, Sehgal PB, Redington M, Kilian PL, Chanock RM, Prince GA. A mouse model for investigating the molecular pathogenesis of adenovirus pneumonia. Proc Natl Acad Sci U S A. 1991;88(5):1651–1655. doi: 10.1073/pnas.88.5.1651. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Gitlin L, Barchet W, Gilfillan S, Cella M, Beutler B, Flavell RA, Diamond MS, Colonna M. Essential role of mda-5 in type I IFN responses to polyriboinosinic:polyribocytidylic acid and encephalomyocarditis picornavirus. Proc Natl Acad Sci U S A. 2006;103(22):8459–8464. doi: 10.1073/pnas.0603082103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Granberg F, Svensson C, Pettersson U, Zhao H. Adenovirus-induced alterations in host cell gene expression prior to the onset of viral gene expression. Virology. 2006;353(1):1–5. doi: 10.1016/j.virol.2006.06.019. [DOI] [PubMed] [Google Scholar]
  23. Haberberger TC, Kupfer K, Murphy JE. Profiling of genes which are differentially expressed in mouse liver in response to adenoviral vectors and delivered genes. Gene Ther. 2000;7(11):903–909. doi: 10.1038/sj.gt.3301181. [DOI] [PubMed] [Google Scholar]
  24. Harrod KS, Trapnell BC, Otake K, Korfhagen TR, Whitsett JA. SP-A enhances viral clearance and inhibits inflammation after pulmonary adenoviral infection. Am J Physiol. 1999;277(3 Pt 1):L580–L588. doi: 10.1152/ajplung.1999.277.3.L580. [DOI] [PubMed] [Google Scholar]
  25. Hartman ZC, Black EP, Amalfitano A. Adenoviral infection induces a multi-faceted innate cellular immune response that is mediated by the toll-like receptor pathway in A549 cells. Virology. 2007a;358(2):357–372. doi: 10.1016/j.virol.2006.08.041. [DOI] [PubMed] [Google Scholar]
  26. Hartman ZC, Kiang A, Everett RS, Serra D, Yang XY, Clay TM, Amalfitano A. Adenovirus infection triggers a rapid, MyD88-regulated transcriptome response critical to acute-phase and adaptive immune responses in vivo . J Virol. 2007b;81(4):1796–1812. doi: 10.1128/JVI.01936-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Haynes LM, Moore DD, Kurt-Jones EA, Finberg RW, Anderson LJ, Tripp RA. Involvement of toll-like receptor 4 in innate immunity to respiratory syncytial virus. J Virol. 2001;75(22):10730–10737. doi: 10.1128/JVI.75.22.10730-10737.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Hensley SE, Cun AS, Giles-Davis W, Li Y, Xiang Z, Lasaro MO, Williams BR, Silverman RH, Ertl HC. Type I interferon inhibits antibody responses induced by a chimpanzee adenovirus vector. Mol Ther. 2007;15(2):393–403. doi: 10.1038/sj.mt.6300024. [DOI] [PubMed] [Google Scholar]
  29. Hensley SE, Giles-Davis W, McCoy KC, Weninger W, Ertl HC. Dendritic cell maturation, but not CD8+ T cell induction, is dependent on type I IFN signaling during vaccination with adenovirus vectors. J Immunol. 2005;175(9):6032–41. doi: 10.4049/jimmunol.175.9.6032. [DOI] [PubMed] [Google Scholar]
  30. Higginbotham JN, Seth P, Blaese RM, Ramsey WJ. The release of inflammatory cytokines from human peripheral blood mononuclear cells in vitro following exposure to adenovirus variants and capsid. Hum Gene Ther. 2002;13(1):129–141. doi: 10.1089/10430340152712683. [DOI] [PubMed] [Google Scholar]
  31. Hirschowitz EA, Weaver JD, Hidalgo GE, Doherty DE. Murine dendritic cells infected with adenovirus vectors show signs of activation. Gene Ther. 2000;7(13):1112–1120. doi: 10.1038/sj.gt.3301210. [DOI] [PubMed] [Google Scholar]
  32. Hoffmann JA, Kafatos FC, Janeway CA, Ezekowitz RA. Phylogenetic perspectives in innate immunity. Science. 1999;284(5418):1313–1318. doi: 10.1126/science.284.5418.1313. [DOI] [PubMed] [Google Scholar]
  33. Iacobelli-Martinez M, Nemerow GR. Preferential activation of Toll-like receptor nine by CD46-utilizing adenoviruses. J Virol. 2007;81(3):1305–1312. doi: 10.1128/JVI.01926-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Irizarry RA, Warren D, Spencer F, Kim IF, Biswal S, Frank BC, Gabrielson E, Garcia JG, Geoghegan J, Germino G, Griffin C, Hilmer SC, Hoffman E, Jedlicka AE, Kawasaki E, Martinez-Murillo F, Morsberger L, Lee H, Petersen D, Quackenbush J, Scott A, Wilson M, Yang Y, Ye SQ, Yu W. Multiple-laboratory comparison of microarray platforms. Nat Methods. 2005;2(5):345–350. doi: 10.1038/nmeth756. [DOI] [PubMed] [Google Scholar]
  35. Ishii KJ, Coban C, Kato H, Takahashi K, Torii Y, Takeshita F, Ludwig H, Sutter G, Suzuki K, Hemmi H, Sato S, Yamamoto M, Uematsu S, Kawai T, Takeuchi O, Akira S. A Toll-like receptor-independent antiviral response induced by double-stranded B-form DNA. Nat Immunol. 2006;7(1):40–48. doi: 10.1038/ni1282. [DOI] [PubMed] [Google Scholar]
  36. Jiang H, Wang Z, Serra D, Frank MM, Amalfitano A. Recombinant adenovirus vectors activate the alternative complement pathway, leading to the binding of human complement protein C3 independent of anti-ad antibodies. Mol Ther. 2004;10(6):1140–1142. doi: 10.1016/j.ymthe.2004.08.015. [DOI] [PubMed] [Google Scholar]
  37. Jooss K, Chirmule N. Immunity to adenovirus and adeno-associated viral vectors: implications for gene therapy. Gene Ther. 2003;10(11):955–63. doi: 10.1038/sj.gt.3302037. [DOI] [PubMed] [Google Scholar]
  38. Kato H, Sato S, Yoneyama M, Yamamoto M, Uematsu S, Matsui K, Tsujimura T, Takeda K, Fujita T, Takeuchi O, Akira S. Cell type-specific involvement of RIG-I in antiviral response. Immunity. 2005;23(1):19–28. doi: 10.1016/j.immuni.2005.04.010. [DOI] [PubMed] [Google Scholar]
  39. Kato H, Takeuchi O, Sato S, Yoneyama M, Yamamoto M, Matsui K, Uematsu S, Jung A, Kawai T, Ishii KJ, Yamaguchi O, Otsu K, Tsujimura T, Koh CS, Reis e, Sousa C, Matsuura Y, Fujita T, Akira S. Differential roles of MDA5 and RIG-I helicases in the recognition of RNA viruses. Nature. 2006;441(7089):101–105. doi: 10.1038/nature04734. [DOI] [PubMed] [Google Scholar]
  40. Kiang A, Hartman ZC, Everett RS, Serra D, Jiang H, Frank MM, Amalfitano A. Multiple innate inflammatory responses induced after systemic adenovirus vector delivery depend on a functional complement system. Mol Ther. 2006;14(4):588–598. doi: 10.1016/j.ymthe.2006.03.024. [DOI] [PubMed] [Google Scholar]
  41. Koizumi N, Yamaguchi T, Kawabata K, Sakurai F, Sasaki T, Watanabe Y, Hayakawa T, Mizuguchi H. Fiber-modified adenovirus vectors decrease liver toxicity through reduced IL-6 production. J Immunol. 2007;178(3):1767–1773. doi: 10.4049/jimmunol.178.3.1767. [DOI] [PubMed] [Google Scholar]
  42. Korst RJ, Mahtabifard A, Yamada R, Crystal RG. Effect of adenovirus gene transfer vectors on the immunologic functions of mouse dendritic cells. Mol Ther. 2002;5(3):307–315. doi: 10.1006/mthe.2002.0538. [DOI] [PubMed] [Google Scholar]
  43. Krieg AM, Wu T, Weeratna R, Efler SM, Love-Homan L, Yang L, Yi AK, Short D, Davis HL. Sequence motifs in adenoviral DNA block immune activation by stimulatory CpG motifs. Proc Natl Acad Sci U S A. 1998;95(21):12631–12636. doi: 10.1073/pnas.95.21.12631. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Kritz AB, Nicol CG, Dishart KL, Nelson R, Holbeck S, Von Seggern DJ, Work LM, McVey JH, Nicklin SA, Baker AH. Adenovirus 5 fibers mutated at the putative HSPG-binding site show restricted retargeting with targeting peptides in the HI loop. Mol Ther. 2007;15(4):741–749. doi: 10.1038/sj.mt.6300094. [DOI] [PubMed] [Google Scholar]
  45. Kumar H, Kawai T, Kato H, Sato S, Takahashi K, Coban C, Yamamoto M, Uematsu S, Ishii KJ, Takeuchi O, Akira S. Essential role of IPS-1 in innate immune responses against RNA viruses. J Exp Med. 2006;203(7):1795–1803. doi: 10.1084/jem.20060792. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Kurt-Jones EA, Sandor F, Ortiz Y, Bowen GN, Counter SL, Wang TC, Finberg RW. Use of murine embryonic fibroblasts to define Toll-like receptor activation and specificity. J Endotoxin Res. 2004;10(6):419–424. doi: 10.1179/096805104225006516. [DOI] [PubMed] [Google Scholar]
  47. Lee MS, Kim YJ. Pattern-recognition receptor signaling initiated from extracellular, membrane, and cytoplasmic space. Mol Cells. 2007a;23(1):1–10. [PubMed] [Google Scholar]
  48. Lee MS, Kim YJ. Signaling pathways downstream of pattern-recognition receptors and their cross talk. Annu Rev Biochem. 2007b;76:447–480. doi: 10.1146/annurev.biochem.76.060605.122847. [DOI] [PubMed] [Google Scholar]
  49. Lieber A, He CY, Meuse L, Schowalter D, Kirillova I, Winther B, Kay MA. The role of Kupffer cell activation and viral gene expression in early liver toxicity after infusion of recombinant adenovirus vectors. J Virol. 1997;71(11):8798–8807. doi: 10.1128/jvi.71.11.8798-8807.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Liu Q, Muruve DA. Molecular basis of the inflammatory response to adenovirus vectors. Gene Ther. 2003;10(11):935–940. doi: 10.1038/sj.gt.3302036. [DOI] [PubMed] [Google Scholar]
  51. Liu Q, White LR, Clark SA, Heffner DJ, Winston BW, Tibbles LA, Muruve DA. Akt/protein kinase B activation by adenovirus vectors contributes to NFkappaB-dependent CXCL10 expression. J Virol. 2005;79(23):14507–14515. doi: 10.1128/JVI.79.23.14507-14515.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Liu Q, Zaiss AK, Colarusso P, Patel K, Haljan G, Wickham TJ, Muruve DA. The role of capsid-endothelial interactions in the innate immune response to adenovirus vectors. Hum Gene Ther. 2003;14(7):627–643. doi: 10.1089/104303403321618146. [DOI] [PubMed] [Google Scholar]
  53. Lopez CB, Moltedo B, Alexopoulou L, Bonifaz L, Flavell RA, Moran TM. TLR-independent induction of dendritic cell maturation and adaptive immunity by negative-strand RNA viruses. J Immunol. 2004;173(11):6882–6889. doi: 10.4049/jimmunol.173.11.6882. [DOI] [PubMed] [Google Scholar]
  54. Loser P, Jennings GS, Strauss M, Sandig V. Reactivation of the previously silenced cytomegalovirus major immediate-early promoter in the mouse liver: involvement of NFkappaB. J Virol. 1998;72(1):180–190. doi: 10.1128/jvi.72.1.180-190.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Lund J, Sato A, Akira S, Medzhitov R, Iwasaki A. Toll-like receptor 9-mediated recognition of Herpes simplex virus-2 by plasmacytoid dendritic cells. J Exp Med. 2003;198(3):513–20. doi: 10.1084/jem.20030162. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Lyakh LA, Koski GK, Young HA, Spence SE, Cohen PA, Rice NR. Adenovirus type 5 vectors induce dendritic cell differentiation in human CD14(+) monocytes cultured under serum-free conditions. Blood. 2002;99(2):600–608. doi: 10.1182/blood.v99.2.600. [DOI] [PubMed] [Google Scholar]
  57. Malkevitch N, Patterson LJ, Aldrich K, Richardson E, Alvord WG, Robert-Guroff M. A replication competent adenovirus 5 host range mutant-simian immunodeficiency virus (SIV) recombinant priming/subunit protein boosting vaccine regimen induces broad, persistent SIV-specific cellular immunity to dominant and subdominant epitopes in Mamu-A*01 rhesus macaques. J Immunol. 2003;170(8):4281–4289. doi: 10.4049/jimmunol.170.8.4281. [DOI] [PubMed] [Google Scholar]
  58. Martin K, Brie A, Saulnier P, Perricaudet M, Yeh P, Vigne E. Simultaneous CAR- and alpha V integrin-binding ablation fails to reduce Ad5 liver tropism. Mol Ther. 2003;8(3):485–494. doi: 10.1016/s1525-0016(03)00182-5. [DOI] [PubMed] [Google Scholar]
  59. Martina Y, Avitabile D, Piersanti S, Cherubini G, Saggio I. Different modulation of cellular transcription by adenovirus 5, DeltaE1/E3 adenovirus and helper-dependent vectors. Virus Res. 2007 doi: 10.1016/j.virusres.2007.05.019. [DOI] [PubMed] [Google Scholar]
  60. Means TK, Hayashi F, Smith KD, Aderem A, Luster AD. The Toll-like receptor 5 stimulus bacterial flagellin induces maturation and chemokine production in human dendritic cells. J Immunol. 2003;170(10):5165–5175. doi: 10.4049/jimmunol.170.10.5165. [DOI] [PubMed] [Google Scholar]
  61. Medzhitov R, Janeway CA., Jr Decoding the patterns of self and nonself by the innate immune system. Science. 2002;296(5566):298–300. doi: 10.1126/science.1068883. [DOI] [PubMed] [Google Scholar]
  62. Miller G, Lahrs S, Shah AB, DeMatteo RP. Optimization of dendritic cell maturation and gene transfer by recombinant adenovirus. Cancer Immunol Immunother. 2003;52(6):347–358. doi: 10.1007/s00262-003-0379-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Montoya M, Schiavoni G, Mattei F, Gresser I, Belardelli F, Borrow P, Tough DF. Type I interferons produced by dendritic cells promote their phenotypic and functional activation. Blood. 2002;99(9):3263–3271. doi: 10.1182/blood.v99.9.3263. [DOI] [PubMed] [Google Scholar]
  64. Morral N, O'Neal WK, Rice K, Leland MM, Piedra PA, Aguilar-Cordova E, Carey KD, Beaudet AL, Langston C. Lethal toxicity, severe endothelial injury, and a threshold effect with high doses of an adenoviral vector in baboons. Hum Gene Ther. 2002;13(1):143–154. doi: 10.1089/10430340152712692. [DOI] [PubMed] [Google Scholar]
  65. Muruve DA, Barnes MJ, Stillman IE, Libermann TA. Adenoviral gene therapy leads to rapid induction of multiple chemokines and acute neutrophil-dependent hepatic injury in vivo . Hum Gene Ther. 1999;10(6):965–976. doi: 10.1089/10430349950018364. [DOI] [PubMed] [Google Scholar]
  66. Ni S, Bernt K, Gaggar A, Li ZY, Kiem HP, Lieber A. Evaluation of biodistribution and safety of adenovirus vectors containing group B fibers after intravenous injection into baboons. Hum Gene Ther. 2005;16(6):664–677. doi: 10.1089/hum.2005.16.664. [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Nicklin SA, Wu E, Nemerow GR, Baker AH. The influence of adenovirus fiber structure and function on vector development for gene therapy. Mol Ther. 2005;12(3):384–393. doi: 10.1016/j.ymthe.2005.05.008. [DOI] [PubMed] [Google Scholar]
  68. Nicol CG, Graham D, Miller WH, White SJ, Smith TA, Nicklin SA, Stevenson SC, Baker AH. Effect of adenovirus serotype 5 fiber and penton modifications on in vivo tropism in rats. Mol Ther. 2004;10(2):344–354. doi: 10.1016/j.ymthe.2004.05.020. [DOI] [PubMed] [Google Scholar]
  69. Nociari M, Ocheretina O, Schoggins JW, Falck-Pedersen E. Sensing infection by adenovirus: Toll-like receptor-independent viral DNA recognition signals activation of the interferon regulatory factor 3 master regulator. J Virol. 2007;81(8):4145–4157. doi: 10.1128/JVI.02685-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. O'Riordan CR, Lachapelle A, Delgado C, Parkes V, Wadsworth SC, Smith AE, Francis GE. PEGylation of adenovirus with retention of infectivity and protection from neutralizing antibody in vitro and in vivo . Hum Gene Ther. 1999;10(8):1349–1358. doi: 10.1089/10430349950018021. [DOI] [PubMed] [Google Scholar]
  71. Parks RJ, Chen L, Anton M, Sankar U, Rudnicki MA, Graham FL. A helper-dependent adenovirus vector system: removal of helper virus by Cre-mediated excision of the viral packaging signal. Proc Natl Acad Sci U S A. 1996;93(24):13565–13570. doi: 10.1073/pnas.93.24.13565. [DOI] [PMC free article] [PubMed] [Google Scholar]
  72. Philpott NJ, Nociari M, Elkon KB, Falck-Pedersen E. Adenovirus-induced maturation of dendritic cells through a PI3 kinase-mediated TNF-alpha induction pathway. Proc Natl Acad Sci U S A. 2004;101(16):6200–6205. doi: 10.1073/pnas.0308368101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Poltorak A, He X, Smirnova I, Liu MY, Van Huffel C, Du X, Birdwell D, Alejos E, Silva M, Galanos C, Freudenberg M, Ricciardi-Castagnoli P, Layton B, Beutler B. Defective LPS signaling in C3H/HeJ and C57BL/10ScCr mice: mutations in Tlr4 gene. Science. 1998;282(5396):2085–2088. doi: 10.1126/science.282.5396.2085. [DOI] [PubMed] [Google Scholar]
  74. Rassa JC, Meyers JL, Zhang Y, Kudaravalli R, Ross SR. Murine retroviruses activate B cells via interaction with toll-like receptor 4. Proc Natl Acad Sci U S A. 2002;99(4):2281–2286. doi: 10.1073/pnas.042355399. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Ross GF, Morris RE, Ciraolo G, Huelsman K, Bruno M, Whitsett JA, Baatz JE, Korfhagen TR. Surfactant protein A-polylysine conjugates for delivery of DNA to airway cells in culture. Hum Gene Ther. 1995;6(1):31–40. doi: 10.1089/hum.1995.6.1-31. [DOI] [PubMed] [Google Scholar]
  76. Santra S, Seaman MS, Xu L, Barouch DH, Lord CI, Lifton MA, Gorgone DA, Beaudry KR, Svehla K, Welcher B, Chakrabarti BK, Huang Y, Yang ZY, Mascola JR, Nabel GJ, Letvin NL. Replication-defective adenovirus serotype 5 vectors elicit durable cellular and humoral immune responses in nonhuman primates. J Virol. 2005;79(10):6516–6522. doi: 10.1128/JVI.79.10.6516-6522.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Schiedner G, Hertel S, Johnston M, Dries V, van Rooijen N, Kochanek S. Selective depletion or blockade of Kupffer cells leads to enhanced and prolonged hepatic transgene expression using high-capacity adenoviral vectors. Mol Ther. 2003;7(1):35–43. doi: 10.1016/s1525-0016(02)00017-5. [DOI] [PubMed] [Google Scholar]
  78. Scibetta AG, Copier J, Barrett A, Chaplin T, Taylor-Papadimitriou J. Gene expression changes induced by a recombinant E1-/E3- adenovirus type 5 vector in human mammary epithelial cells. Intervirology. 2005;48(6):350–361. doi: 10.1159/000086062. [DOI] [PubMed] [Google Scholar]
  79. Shayakhmetov DM, Gaggar A, Ni S, Li ZY, Lieber A. Adenovirus binding to blood factors results in liver cell infection and hepatotoxicity. J Virol. 2005a;79(12):7478–7491. doi: 10.1128/JVI.79.12.7478-7491.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. Shayakhmetov DM, Li ZY, Ni S, Lieber A. Analysis of adenovirus sequestration in the liver, transduction of hepatic cells, and innate toxicity after injection of fiber-modified vectors. J Virol. 2004;78(10):5368–5381. doi: 10.1128/JVI.78.10.5368-5381.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  81. Shayakhmetov DM, Li ZY, Ni S, Lieber A. Interference with the IL-1-signaling pathway improves the toxicity profile of systemically applied adenovirus vectors. J Immunol. 2005b;174(11):7310–7319. doi: 10.4049/jimmunol.174.11.7310. [DOI] [PubMed] [Google Scholar]
  82. Shi L, Tong W, Fang H, Scherf U, Han J, Puri RK, Frueh FW, Goodsaid FM, Guo L, Su Z, Han T, Fuscoe JC, Xu ZA, Patterson TA, Hong H, Xie Q, Perkins RG, Chen JJ, Casciano DA. Cross-platform comparability of microarray technology: intra-platform consistency and appropriate data analysis procedures are essential. BMC Bioinformatics. 2005;(6 Suppl 2):S12. doi: 10.1186/1471-2105-6-S2-S12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  83. Smith TA, Idamakanti N, Marshall-Neff J, Rollence ML, Wright P, Kaloss M, King L, Mech C, Dinges L, Iverson WO, Sherer AD, Markovits JE, Lyons RM, Kaleko M, Stevenson SC. Receptor interactions involved in adenoviral-mediated gene delivery after systemic administration in non-human primates. Hum Gene Ther. 2003a;14(17):1595–1604. doi: 10.1089/104303403322542248. [DOI] [PubMed] [Google Scholar]
  84. Smith TA, Idamakanti N, Rollence ML, Marshall-Neff J, Kim J, Mulgrew K, Nemerow GR, Kaleko M, Stevenson SC. Adenovirus serotype 5 fiber shaft influences in vivo gene transfer in mice. Hum Gene Ther. 2003b;14(8):777–787. doi: 10.1089/104303403765255165. [DOI] [PubMed] [Google Scholar]
  85. Snoeys J, Mertens G, Lievens J, van Berkel T, Collen D, Biessen EA, De Geest B. Lipid emulsions potently increase transgene expression in hepatocytes after adenoviral transfer. Mol Ther. 2006;13(1):98–107. doi: 10.1016/j.ymthe.2005.06.477. [DOI] [PubMed] [Google Scholar]
  86. Stilwell JL, Samulski RJ. Role of viral vectors and virion shells in cellular gene expression. Mol Ther. 2004;9(3):337–346. doi: 10.1016/j.ymthe.2003.11.007. [DOI] [PubMed] [Google Scholar]
  87. Stone D, Liu Y, Shayakhmetov D, Li ZY, Ni S, Lieber A. Adenovirus-platelet interaction in blood causes virus sequestration to the reticuloendothelial system of the liver. J Virol. 2007;81(9):4866–4871. doi: 10.1128/JVI.02819-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  88. Sun Q, Sun L, Liu HH, Chen X, Seth RB, Forman J, Chen ZJ. The specific and essential role of MAVS in antiviral innate immune responses. Immunity. 2006;24(5):633–642. doi: 10.1016/j.immuni.2006.04.004. [DOI] [PubMed] [Google Scholar]
  89. Tamanini A, Nicolis E, Bonizzato A, Bezzerri V, Melotti P, Assael BM, Cabrini G. Interaction of adenovirus type 5 fiber with the coxsackievirus and adenovirus receptor activates inflammatory response in human respiratory cells. J Virol. 2006;80(22):11241–11254. doi: 10.1128/JVI.00721-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. Tan PK, Downey TJ, Spitznagel EL, Jr, Xu P, Fu D, Dimitrov DS, Lempicki RA, Raaka BM, Cam MC. Evaluation of gene expression measurements from commercial microarray platforms. Nucleic Acids Res. 2003;31(19):5676–5684. doi: 10.1093/nar/gkg763. [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. Tao N, Gao GP, Parr M, Johnston J, Baradet T, Wilson JM, Barsoum J, Fawell SE. Sequestration of adenoviral vector by Kupffer cells leads to a nonlinear dose response of transduction in liver. Mol Ther. 2001;3(1):28–35. doi: 10.1006/mthe.2000.0227. [DOI] [PubMed] [Google Scholar]
  92. Tarodi B, Metz DH, Douglas A, Russell WC. A Study of events in chick cells infected with human adenovirus type 5 and their relationship to the induction of interferon. J Gen Virol. 1977;36(3):425–436. doi: 10.1099/0022-1317-36-3-425. [DOI] [PubMed] [Google Scholar]
  93. Thorne PS, McCray PB, Howe TS, O'Neill MA. Early-onset inflammatory responses in vivo to adenoviral vectors in the presence or absence of lipopolysaccharide-induced inflammation. Am J Respir Cell Mol Biol. 1999;20(6):1155–1164. doi: 10.1165/ajrcmb.20.6.3632. [DOI] [PubMed] [Google Scholar]
  94. Tibbles LA, Spurrell JC, Bowen GP, Liu Q, Lam M, Zaiss AK, Robbins SM, Hollenberg MD, Wickham TJ, Muruve DA. Activation of p38 and ERK signaling during adenovirus vector cell entry lead to expression of the C-X-C chemokine IP-10. J Virol. 2002;76(4):1559–1568. doi: 10.1128/JVI.76.4.1559-1568.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  95. Trevejo JM, Marino MW, Philpott N, Josien R, Richards EC, Elkon KB, Falck-Pedersen E. TNF-alpha -dependent maturation of local dendritic cells is critical for activating the adaptive immune response to virus infection. Proc Natl Acad Sci U S A. 2001;98(21):12162–12167. doi: 10.1073/pnas.211423598. [DOI] [PMC free article] [PubMed] [Google Scholar]
  96. Ustacelebi S, Williams JF. Temperature-sensitive mutants of adenovirus defective in interferon induction at non-permissive temperature. Nature. 1972;235(5332):52–53. doi: 10.1038/235052a0. [DOI] [PubMed] [Google Scholar]
  97. Varnavski AN, Calcedo R, Bove M, Gao G, Wilson JM. Evaluation of toxicity from high-dose systemic administration of recombinant adenovirus vector in vector-naive and pre-immunized mice. Gene Ther. 2005;12(5):427–436. doi: 10.1038/sj.gt.3302347. [DOI] [PubMed] [Google Scholar]
  98. Varnavski AN, Schlienger K, Bergelson JM, Gao GP, Wilson JM. Efficient transduction of human monocyte-derived dendritic cells by chimpanzee-derived adenoviral vector. Hum Gene Ther. 2003;14(6):533–544. doi: 10.1089/104303403764539323. [DOI] [PubMed] [Google Scholar]
  99. Varnavski AN, Zhang Y, Schnell M, Tazelaar J, Louboutin JP, Yu QC, Bagg A, Gao GP, Wilson JM. Preexisting immunity to adenovirus in rhesus monkeys fails to prevent vector-induced toxicity. J Virol. 2002;76(11):5711–5719. doi: 10.1128/JVI.76.11.5711-5719.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Volk AL, Rivera AA, Page GP, Salazar-Gonzalez JF, Nettelbeck DM, Matthews QL, Curiel DT. Employment of microarray analysis to characterize biologic differences associated with tropism-modified adenoviral vectors: utilization of non-native cellular entry pathways. Cancer Gene Ther. 2005;12(2):162–174. doi: 10.1038/sj.cgt.7700776. [DOI] [PubMed] [Google Scholar]
  101. Woo Y, Affourtit J, Daigle S, Viale A, Johnson K, Naggert J, Churchill G. A comparison of cDNA, oligonucleotide, and Affymetrix GeneChip gene expression microarray platforms. J Biomol Tech. 2004;15(4):276–284. [PMC free article] [PubMed] [Google Scholar]
  102. Worgall S, Wolff G, Falck-Pedersen E, Crystal RG. Innate immune mechanisms dominate elimination of adenoviral vectors following in vivo administration. Hum Gene Ther. 1997;8(1):37–44. doi: 10.1089/hum.1997.8.1-37. [DOI] [PubMed] [Google Scholar]
  103. Xu LG, Wang YY, Han KJ, Li LY, Zhai Z, Shu HB. VISA is an adapter protein required for virus-triggered IFN-beta signaling. Mol Cell. 2005;19(6):727–740. doi: 10.1016/j.molcel.2005.08.014. [DOI] [PubMed] [Google Scholar]
  104. Yang R, Murillo FM, Cui H, Blosser R, Uematsu S, Takeda K, Akira S, Viscidi RP, Roden RB. Papillomavirus-like particles stimulate murine bone marrow-derived dendritic cells to produce alpha interferon and Th1 immune responses via MyD88. J Virol. 2004;78(20):11152–11160. doi: 10.1128/JVI.78.20.11152-11160.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  105. Yauk CL, Berndt ML, Williams A, Douglas GR. Comprehensive comparison of six microarray technologies. Nucleic Acids Res. 2004;32(15):e124. doi: 10.1093/nar/gnh123. [DOI] [PMC free article] [PubMed] [Google Scholar]
  106. Zaiss AK, Liu Q, Bowen GP, Wong NC, Bartlett JS, Muruve DA. Differential activation of innate immune responses by adenovirus and adeno-associated virus vectors. J Virol. 2002;76(9):4580–4590. doi: 10.1128/JVI.76.9.4580-4590.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  107. Zhu J, Huang X, Yang Y. Innate immune response to adenoviral vectors is mediated by both Toll-like receptor-dependent and -independent pathways. J Virol. 2007;81(7):3170–3180. doi: 10.1128/JVI.02192-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  108. Ziegler RJ, Li C, Cherry M, Zhu Y, Hempel D, van Rooijen N, Ioannou YA, Desnick RJ, Goldberg MA, Yew NS, Cheng SH. Correction of the nonlinear dose response improves the viability of adenoviral vectors for gene therapy of Fabry disease. Hum Gene Ther. 2002;13(8):935–945. doi: 10.1089/10430340252939041. [DOI] [PubMed] [Google Scholar]
  109. Zsengeller Z, Otake K, Hossain SA, Berclaz PY, Trapnell BC. Internalization of adenovirus by alveolar macrophages initiates early proinflammatory signaling during acute respiratory tract infection. J Virol. 2000;74(20):9655–9667. doi: 10.1128/jvi.74.20.9655-9667.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  110. Zsengeller ZK, Wert SE, Bachurski CJ, Kirwin KL, Trapnell BC, Whitsett JA. Recombinant adenoviral vector disrupts surfactant homeostasis in mouse lung. Hum Gene Ther. 1997;8(11):1331–1344. doi: 10.1089/hum.1997.8.11-1331. [DOI] [PubMed] [Google Scholar]

RESOURCES