Next Article in Journal
Avian Influenza A Virus Associations in Wild, Terrestrial Mammals: A Review of Potential Synanthropic Vectors to Poultry Facilities
Next Article in Special Issue
Adenovirus and the Cornea: More Than Meets the Eye
Previous Article in Journal
Recombinant Lactococcus Expressing a Novel Variant of Infectious Bursal Disease Virus VP2 Protein Can Induce Unique Specific Neutralizing Antibodies in Chickens and Provide Complete Protection
Previous Article in Special Issue
Role of CCCH-Type Zinc Finger Proteins in Human Adenovirus Infections
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Adenovirus—Extracellular Protein Interactions and Their Impact on Innate Immune Responses by Human Mononuclear Phagocytes

Institut de Génétique Moléculaire de Montpellier, Université de Montpellier, CNRS (French National Centre for Scientific Research), 34090 Montpellier, France
*
Author to whom correspondence should be addressed.
Submission received: 16 October 2020 / Revised: 23 November 2020 / Accepted: 24 November 2020 / Published: 26 November 2020
(This article belongs to the Special Issue Adenovirus Pathogenesis)

Abstract

:
The aim of this review is to highlight how, in a syngeneic system, human mononuclear phagocytes respond to environments containing human adenovirus (HAdV) and soluble extracellular proteins that influence their innate immune response. Soluble extracellular proteins, including immunoglobulins, blood clotting factors, proteins of the complement system, and/or antimicrobial peptides (AMPs) can exert direct effects by binding to a virus capsid that modifies interactions with pattern recognition receptors and downstream signaling. In addition, the presence, generation, or secretion of extracellular proteins can indirectly influence the response to HAdVs via the activation and recruitment of cells at the site of infection.

In this review, we highlight how human mononuclear phagocytes respond to a milieu containing human adenoviruses (HAdVs) or HAdV-based vectors and an array of soluble extracellular proteins. The response by phagocytes, which include monocytes, macrophages, conventional and plasmacytoid dendritic cells (cDCs and pDCs, respectively), orchestrates long-term immune responses. Adenoviridae is a family of ~150 mDa, 90 nm diameter, nonenveloped, proteinaceous particles containing a linear double-stranded DNA genome of 26–48 kb [1]. Adenoviruses (AdVs) infect mammals, reptiles, birds, fish, and amphibians. Likely due to their impact on human health, the majority of the ~200 officially recognized AdVs has been isolated from humans [2]. In most cases, human AdVs cause self-limiting respiratory, ocular, or gastro-intestinal tract infections in all populations. HAdVs are currently classified into seven species (human adenovirus A to human adenovirus G, HAdV-A to HAdV-G) and ~100 types (identified by serology and/or sequence phylogeny) [1]. Over the last 50 years the vectorization and immunogenicity of HAdVs have been areas of significant interest in the context of vaccines, oncolytics, gene transfer vectors, and morbidity in immune-compromised individuals [3,4,5]. HAdV-based vaccines, including live virus or replication-defective vectors, are used to prevent acute respiratory disease in military recruits [6] (e.g., HAdV type 4 (HAdV-E4) and HAdV-B7), or in preclinical and clinical trials targeting Malaria, Ebola virus, and SARS-CoV-2 [7,8,9,10,11,12].
Contrary to conventional wisdom, very little clinical data exist that demonstrate that HAdVs stimulate the maturation of human mononuclear phagocytes [13]. This is also the case in vitro unless high doses (100,000–200,000 particles/cell) are used [14]. Many environments exist where encounters between HAdVs and mononuclear phagocytes differ, which will differentially orient a de novo adaptive or memory response. In the case of natural infection, HAdVs will first encounter mucosa (respiratory, gastrointestinal, or ocular) surfaces. Whereas, in the case of vaccines, oncolytics, and gene therapy, HAdV vectors could be delivered intradermal, intramuscular, intravenous, orally, to the respiratory tract, in the brain, or directly into tumors [15]. Variations in frequency and activity of immune cell populations present in, or migrating to, different sites of delivery lead to differences in responses. In addition, a fourth dimension—time—needs to be incorporated into the dynamics of how a host responds to these multiparametric encounters.

1. Extracellular Environments

Soluble extracellular proteins englobe a spectrum of compounds that are secreted from cells in their mature form or modified by extracellular processing [16]. In plasma, albumin (54%), globulins (37%) and fibrinogen (7%) represent the major constituents [17]. Antibodies (Abs) are the first globulins to come to mind when considering a response to viruses. Although coagulation factors, complement components and antimicrobial peptides (or proteins) (AMPs) make up a small proportion of serum components, their impact on innate and adaptive immune responses is considerable. AMPs are secreted by epithelial cells and/or immune cells; Abs can be found in the systemic circulation or in the mucosa, complement components and coagulation factors are in the plasma [18].

2. Immunoglobulins

In the context of interactions with HAdVs, IgMs (naturally occurring, typically pentameric, high avidity and the first immunoglobulin to encounter pathogens), IgGs (the most prevalent antigen-specific immunoglobulins induced rapidly after antigen encounters), and IgAs (found preferentially in mucosal tissues) are most relevant. Numerous studies have demonstrated that antibody–antigen complexes are efficient stimulators of DC maturation and inflammation [19]. However, the scenario of intravenous injection of large doses of HAdV vectors in an immunologically naïve patient is a situation that would be rarely encountered (in contrast to assays in mice). More than 90% of the world population have encountered at least one, and likely several HAdVs types, before reaching the age of 5 years old [20]. Therefore, it is not surprising that HAdV-ICs (immune complexes) are found in some individuals with severe HAdV disease [20,21]. An immunology paradigm, based on the ability of some Abs to neutralize infection of some cells, predicts that Abs protect us from pathogens. By contrast, antibody-dependent enhanced infection can make some viral encounters worse for the host. A recent example of this is how some B-cell responses to flavivirus infections increase the pathogenicity of other members of this family [22,23]. We and others have repeatedly shown that while an Ab can neutralize HAdV infection of epithelial-like cells; this does not correlate with neutralization on cells expressing Fcγ receptors (FcγR) [24,25,26,27].
The mechanistic difference here lies in part in receptor use and intracellular processing: immunoglobulin neutralization of HAdVs by epithelial cells occurs primarily by inhibiting interactions with the virus receptor, or by post-internalization mechanisms that prevent capsid disassembly and degradation by lysosomes. At least in the case of HAdV-C5, the primary receptor switches from coxsackievirus adenovirus receptor (CAR) for epithelial cells to DC-SIGN (dendritic cell-specific intercellular adhesion molecule-3-grabbing non-integrin also is a C-type lectin receptor) for DCs and macrophages [28] (this is unlikely the case for HAdVs that use CD46 (some HAdV-Bs) or sialic acid-modified proteins motifs (some HAdV-Ds)). We are unaware of studies that map the region of HAdV-C5 that interacts with DC-SIGN and whether naturally-occurring Abs block this interaction.
Once in the cytoplasm, HAdVs encounter a vastly different environment in epithelial cells versus phagocytes. Notably, DCs slowly degrade proteins to allow cross-presentation via the MHC I complex [29]. Therefore, whether a HAdV is taken up by direct interaction with a receptor, or by redirecting to FcγRs expressed by phagocytes, dictates how the immune response develops. Differential internalization pathways are supported by the observation that HAdV-IC are found clustered into large aggregates in DCs [24], which may facilitate the escape from lysosome-mediated degradation by some particles. Perreau et al. showed that HAdV-IC s bind FcγRs and are internalized by human DCs [13]. Once HAdV-ICs access the cytosol they may also be engaged by tripartite motif-containing protein 21 (TRIM21) via the Fc portion of the Abs [30]. TRIM21 may contribute to the efficient cytosolic neutralization of mAb-bound HAdV-C5 in vitro [31] and in vivo (in mice) [32]. Again, in a cellular model, Ewan et al. showed that the more TRIM21 is expressed, the more the virus is neutralized [33]. Eichholz et al. then extended these observations to show that HAdV-ICs induce absent in melanoma 2 (AIM2) engagement and pyroptotic DC death. Pyroptosis, an inflammatory cell death that occurs in some mononuclear phagocytes following activation of inflammasome formation, leads to the release of pro-inflammatory cytokines such as TNF, IL-6 and IL-1β [24]. Under natural HAdV infection conditions, this may be a pro-host response. However, in the context of using a HAdV-C5 vector as a vaccine against HIV infections [34,35], pyroptosis may have set up a dynamic situation where HAdV-specific T cells were induced to proliferate, home to mucosal sites, and become prime targets for HIV infection [36,37].
Although the analyses are low in power and irreproducible due to sample availability, antibody-dependent enhanced infection of mononuclear phagocytes might have been involved in the death of a patient going under HAdV-C5-mediated gene therapy for ornithine transcarbamylase deficiency in the 1990’s [38]. These data challenge the safety of HAdV-based therapy in patients with preexisting anti-HAdV immunity [12]. The impact of anti-HAdV Abs also took another twist when Tran et al. [27] showed that the environment created by DCs undergoing HAdV-IC-induced pyroptosis induces bystander DCs to become tolerogenic. These antigen-loaded tolerogenic DCs can drive naïve T cells to mature into HAdV-specific TREGs [27]. What role HAdV-specific TREGs play in HAdV infection and disease is an open question because their effect can be extremely plastic. We have hypothesized that they contribute to persistent HAdV infections and the reduction of possible immune-mediated tissue damage. Another possibility is that harboring a robust anti-HAdV response may generate cross-reactive B-cell responses to more virulent pathogens [39].

3. Complement

The complement system is a network of more than 50 soluble and cell-bound molecules, produced by essentially all cells in mammals, that contribute to innate and adaptive immunity [40]. This self-amplifying cascade of messenger and effector molecules detects breaches in tissue homeostasis and pathogens. There are three canonical complement pathways: classical (predominantly Ig-mediated); alternative (interaction of complement proteins with microbial components); and mannose-binding lectin [41]. A noncanonical pathway consists of an Ig-dependent, component 3 (C3)-independent mechanism [42]. Typical complement-induced responses are self-limiting, but uncontrolled amplification contributes to disseminated intravascular coagulation, sepsis, and adult respiratory distress syndrome [43].
The classical and alternative pathways are the most relevant in the context of HAdVs. In the case of the classical pathway, the attachment of Ig to a HAdV capsid induces binding of C1q to initiate the cascade (numerous reviews can be found describing the C1q–C9 cascade and therefore it will not be repeated here [44]). In pioneering studies, Cichon et al. showed that HAdV–C5-ICs can activate complement [45]. Interestingly, they also were the first to show that high concentrations of HAdV–C5 directly activate a complement pathway, but did not characterize the process. Later, it was shown that HAdV–C5 can induce the alternative complement pathway by binding C3 directly [46]. Lastly, a C1q-dependent mechanism can neutralize HAdV–C5 by covalently linking C4b, which then targets the complex for gC1qBP-mediated uptake where HAdV–C5 is degraded in lysosomes [47]. It is also worth noting that there are exceptions: although human sera can contain Abs that bind canine adenovirus type 2 (CAV-2), the classical and alternative complement pathways are not engaged due to a block upstream of the C3 convertase and interplay between the C1q–C1r2–C1s2 complex and C1q inhibitor and CAV-2 [48].
As the focus of this review is to link HAdVs, extracellular proteins, and human phagocytes, we have focused on the cellular events downstream of complement activation. All HAdV-associated complement pathways result in the generation of the anaphylatoxins C3a and C5a. Via their cognate receptors C3aR, and C5aR1, and C5aR2, C3a, and C5a exert regulatory effects on the humoral and cellular arms of innate immunity [49]. For example, C3b induces IL-1 and TNF release [50] and activation of phagocytes [51]. Yet, unexpectedly, using an in vitro model system, Zaiss et al. concluded that complement did not affect HAdV-C5 entry or activation of human macrophages [25]. While possible in mice, it is unknown if HAdV-C5 uptake by human neutrophil is affected by the presence of complement or Abs. Notably, HAdV-C5 and neutrophil interactions were significantly reduced when they were incubated in IgG-depleted or serum containing complement [52].
C5a is a potent inflammatory peptide with a broad spectrum of functions including acting as a chemoattractant for neutrophils, monocytes, and macrophages [53]. In humans, C5aR1 is on the surface of most DC subsets, including monocyte-derived DCs, dermal DCs, and Langerhans cells [54]. The direct effects of C5aR1 engagement leads to the activation of pattern-recognition receptors and FcγRs. Examples of this cross-talk include C3a/C3aR-mediated activation of the inflammasome in human macrophages, monocytes and DCs [55,56], and C5a-mediated suppression of Toll-like receptor (TLR)-induced IL-12 production from macrophages [57], as well as C5a-mediated activation of phagocytes through IC engagement of the inhibitory IgG-binding FcγRIIb with Dectin-1 [58]. C3a and C5a are also involved in the development of Th2 and Th17 immunity through the tweaking of DC activation [59], and the induction of regulatory T cells (TREGs) [60]. Whether complement-linked induction of TREGs synergizes with the HAdV-IC induced tolerogenic DC and TREGs [27] is unknown and worthy of more investigation.
On the flip side, thrombocytopenia and anemia can be associated with some severe cases of adenoviremia. Why is this relevant to complement activation? In a GPIb- and C3-dependent manner, platelets can deliver pathogens to tissue resident DCs and increase the pathogen directed immune response [61]. As mentioned above, the direct binding of C3 to some HAdV capsids could influence the anti-HAdV response when they are found in the systemic circulation (whether due to wild-type infections or delivery for gene therapy). In addition to platelets, some adenovirus types that bind CAR induce hemagglutination of human erythrocytes [62]. Therefore, immune adherence may also contribute to the complement-associated response to HAdVs [63] as C3–HAdV complexes could be delivered to the complement receptor 1 (CR1) clustered on the surface of human erythrocytes [64]. It is worth noting that C5a can also activate the coagulation pathway [65], which brings immunoglobulins, complement, and coagulation factors into the global discussion of an anti-HAdV immune response.

4. Antimicrobial Peptides (AMPs)

AMPs are effector molecules of the innate immune system and can act as an endogenous antimicrobial against a broad array of infectious agents [66,67,68,69]. More than 1300 AMPs have been identified and many are produced by epithelial cells of skin, oral mucosa, and gastrointestinal tract, and by myeloid cells, including neutrophils, which are among the first leukocytes to infiltrate infected and vector-injected tissues [36]. The ubiquitous expression and rapid production of AMPs allow them to act as first line responders to pathogens [69,70]. In addition to responses to pathogens, AMPs can recruit and modulate the activities of neutrophils, monocytes, macrophages, DCs and T cells. AMPs can induce the expression of cytokines, neutralize endotoxin and induce angiogenesis [66,67,68]. Minor variations in the primary sequence of an AMP can induce major changes in their activity and, therefore, syngeneic systems are the gold standard to understand clinically relevant scenarios [68]. Notably, pockets of negatively charged areas in the hexon hypervariable regions of many HAdV types [71,72,73,74,75] favor binding of positively charged AMPs.

4.1. Alarmins

Among the AMPs is a subset dubbed “alarmins”, which include defensins, lactoferrin (Lf), and LL-37. Defensins are ~30 amino acid cationic peptides with a conserved framework of six disulfide-linked cysteines. Two α-defensins, human neutrophil peptides 1 and 2 (HNP-1 and -2) stimulate monocyte, DC and T cell recruitment [76,77,78,79]. The four human β-defensins (hBD1–4) stimulate recruitment of monocytes, DCs, T cells and mast cells [80,81,82]. Both α- and β-defensins also have selective actions: α-defensins induce the recruitment of human CD8+ and CD4+/CD45RA+ naïve T cells, whereas β-defensins induce recruitment of immature DCs and CD4+/CD45RO+ memory T cells [79]. Furthermore, during interaction with TLR-1 and -2 on monocytes and DCs, hBD3 can induce upregulation of costimulatory molecules CD80, CD86 and CD40 [83].
Lf, an atypically large alarmin (80 kDa), is a multifunctional member of the transferrin family that sequesters iron and is found in various secretory fluids, such as milk, saliva, tears, and nasal secretions. Lactoferrin induces DC maturation via DC-SIGN or targeting TLR-2 and -4 and drives Th1 responses [84,85,86,87]. Lf also induces DC maturation. Lactoferricin, a biologically active N-terminal fragment of 49 aa, also exhibits immunomodulatory effects, such as stimulation of monocyte differentiation and release of pro-inflammatory cytokines [88].
LL-37 is a 37 amino acid cationic peptide found in lysosomes of macrophages, polymorphonuclear leukocytes, and keratinocytes. LL-37 serves a critical role against infections by its chemotactic activity [89], influencing phagocytosis [90], TLR activation [91,92,93], DC differentiation, and induced T-cell polarization [94].

4.2. How Alarmins Influence HAdVs

Alarmins influence HAdV infections and immunogenicity via mechanisms that vary in each model system. HNP-1 and human defensin 5 (HD5) impair HAdV-C5, -D12 and -B35 infection of epithelial cells by stabilizing an intrinsically disordered region of the vertex and thus prevent initial stages of viral disassembly post-entry [95,96,97,98]. By contrast, based on protein separation using bronchoalveolar lavage, Skrygan et al. concluded that defensins are not important anti-HAdV-C5 factors during infection of lung epithelium [99]. Recently, Tartaglia et al. showed that HD5 regulates immunogenicity by increasing HAdV-D26 and HAdV-D48 infections in mice (but not HAdV-C5 and HAdV-B35) [100]. It is relevant to note that HD1, 3, and 4 are elevated in nasal aspirates from children with HAdV infections [101]. While the authors assumed that increased levels of defensins should be pro-host, this assumption is challenging to demonstrate in a clinical setting. While not directly related to human phagocytes, tumor-derived HD5 affected oncolytic HAdV-B3 replication during cancer therapy [102]. Interestingly, HD5 inhibition was circumvented by the capacity of HAdV-B3 to produce virus-like particles made up of penton-dodecahedra that sequester and neutralize HD5.
Arnberg and colleagues demonstrated that lactoferrin binds to the negatively-charged hexon hypervariable regions of HAdV-C5, -A31 and -B35 [75], thereby acting as a bridge to increase infection in human epithelial-like cells [103]. Notably, enhanced infection is independent of CAR [104], the primary cellular attachment molecule for HAdV-C5 and -A31 [105]. Lf also mediates increased HAdV-C5 transduction in human monocytes, pDCs, cDCs, and Langerhans cells possibly via the engagement of DC-SIGN [86] and other factors. Interestingly, human Lf has no effect on HAdV-C2 infection of the above phagocytes while bovine Lf increased HAdV transduction [106].
Finally, Uchio et al. found LL-37 also inhibits propagation of HAdV-B3, -D8, -D19a, and -D37, but not -E4 (which is likely of simian origin and now endemic in some military training camps) infection of A549 cells [107]. By contrast, Gordon et al. found that LL-37 inhibits the propagation of HAdV-D19 (but no significant inhibition on HAdV-B3, -C5, or -D8) [108]. These differences may be due to the experimental design with respect to the sequential versus simultaneous addition of virus and LL-37, and the dose of LL-37.

5. Coagulation Factor

The coagulation system is a multiprotein cascade that is involved in blood clot formation [109]. Secreted in the systemic circulation as inactive zymogens, the vitamin K-dependent coagulation factors VII, IX and X all share a common structural N terminal γ-carboxyl glutamic acid domain and a C terminal serine protease domain. It is the N terminus that binds with nanomolar affinity to some of the hexon hypervariable regions [110]. Based on HAdV pathogenicity, it is difficult to explain what selective advantage binding to vitamin K-dependent coagulation factors adds. Numerous studies and reviews address the impact of vitamin K-dependent coagulation factors on HAdV-mediated gene transfer in mice [111,112,113,114], with the global conclusion that vitamin K-dependent coagulation factors act as bridges between the negatively-charged capsid and negatively-charged heparan sulfate proteoglycans. The use of coagulation factors by HAdV may be important for liver tropism following intravenous injection of HAdV-C5 vectors, and during naturally occurring infections. Indeed HAdV-C5 and -A31 require FIX or FX to infect human corneal epithelial (HCE) cells and human lung carcinoma (A549) cells [115]. It is noteworthy that Eichholz et al. found that human FX coupled to HAdV-C5 did not induce greater uptake by, or maturation of, human DCs [116]. Therefore, the mechanism by which some coagulations factors modify the tropism and response to some HAdVs and their vectors needs greater species-specific analyses. That being said, Alba et al. quantified HAdV-C5 and FX-binding-ablated HAdV-C5-mediated gene transfer in a non-human primate and found that HAdV-C5 vectors efficiently mediated gene transfer to the liver, whereas FX-binding-ablated vectors primarily targeted the spleen. Therefore, some data generated using mice may be consistent with the mechanism underlying hepatocyte transduction in humans [117].

6. Perspectives

Whenever possible, we focused on how HAdVs and extracellular proteins impact the innate immune response in human phagocytes. A caveat though is that the use of replication-defective HAdV vectors precludes the analyses of the impact of extracellular proteins on events downstream of entry into the nucleus. When necessary we have introduced data generated in other systems (e.g., rodents) to provide background. For example, HAdVs alone readily induce the maturation of murine DCs, but poorly induce human DC maturation [13,118]. We know that coagulation factor-specific interactions influence HAdV innate immune responses differentially in mouse and in human phagocytes. Murine FX-armored HAdVs stimulated murine macrophages in vitro and in vivo, whereas human FX–HAdV complexes had no effect on human DCs in vitro [116,119]. While bovine Lf increases HAdV-C2 infection of human DCs, human Lf has no effect [106]. Nonetheless, data from rodents can generate avenues to explore in humans. For example, Gounder et al. found that murine α-defensins protect mice from oral infection by mouse adenovirus 1 (MAdV-1). Yet, in spite of the sensitivity of MAdV-1 infection to α-defensin neutralization in cell culture, α-defensin production by Paneth cell had no effect on the kinetics and magnitude of MAdV-1 dissemination to the brain. This may have been due to a delay in the production of neutralizing Abs. Thus, in mice, α-defensins play a role in antiviral immunity that is distinct from their antiviral activity observed in cell culture [120].
The law of parsimony suggests synergistic interactions between Abs, coagulation factors, complement and AMPs in most pathological or vaccine-associated paradigms. Yet, some AMPs can inhibit the complement pathway. Lf, by blocking the formation of C3 convertase [121], and HNP-1, by blocking the C1q hemolytic activity [122], can inhibit the classical complement pathway. Xu et al. demonstrated that FX shields HAdV-C5 and blocks HAdV-dependent complement activation [123]. In each scenario, two parameters are crucial: site and timeline of infection. If HAdVs enter via intranasal, oral, gastrointestinal, ocular, intravenous, intradermal, or muscular tissues, the first cells to be infected, and then those recruited, will differ. For example, some phagocytes (e.g., alveolar macrophages) are recruited to specific organs and not all extracellular proteins will be all present or secreted at the same level.
The timeline of infection and extracellular protein secretion are also crucial parameters. Depending on the extracellular protein first secreted, it could change HAdV interactions with its target cell. For example, if an AMP binds first, does that interaction prevent Ab binding? Do Abs outcompete AMPs because of affinity or avidity? Are there situations where they act synergistically? Unsurprisingly, assays addressing these complex interactions are challenging because of the direct effect of some extracellular proteins on phagocyte homeostasis and recruitment. Yet, these issues are critically important to understand in the scenario where AdV-based vaccines are used [12]. It is possible that some extracellular proteins could be used as adjuvants to increase vaccine efficacy. If an AMP is used as an adjuvant, could the complex increase HAdV uptake by phagocytes? or uptake by other tissue resident cells? There exist multiple modifications of HAdV capsid sequences (or coating the capsid with protective polymers) that reduce binding of extracellular proteins [124,125,126,127]. Testing these HAdV vectors as vaccines could provide valuable information on the complex interactions that alarmins could play. We know very little about what happens in the 48 h post-vaccine injection (predominantly in muscle for vaccines): what cells are recruited and in what ratio? What do they secrete in a host with or without preexisting humoral and cellular immunity? How does this combination of cells and extracellular proteins modify the innate and adaptive immune response? Clearly multiple extracellular proteins are found in the same microenvironment and therefore addressing their impact on HAdV uptake in binary analyses poorly considers HAdV or phagocyte interactions. Add on top tissue undergoing an inflammatory reaction where the extracellular matrix, which is composed of proteins, proteoglycans, and glycoproteins each with varying physical and biochemical properties, is degraded and releases peptides. Matrix metalloproteinases, ADAMTS (a disintegrin and metalloproteinase with thrombospondin motifs), and serine proteinases (plasmin and cathepsin G) are the most common extracellular matrix degrading enzymes [128]. Very little is known about how extracellular matrix degradation modifies HAdV tropism and immunogenicity. Data that may shed light onto this process might be gleaned from oncolytic vectors that express enzymes that degrade the extracellular matrix [129].

Author Contributions

C.C. and E.J.K. wrote the manuscript and have read and agreed to the published version of the manuscript. Both authors have read and agreed to the published version of the manuscript.

Funding

This work benefited from support by the LabEx “EpiGenMed”, an “Investissements d’avenir” program, Université de Montpellier, the Innovative Medicine Initiative EboVac2 (#115861), The Fondation Pierre Fabre, the Agence Nationale de la Recherche (NORAD-19-CE37-0008-04, Goal-14-CE13-0014-03), ERA-Net (E-Rare) Joint Transnational Call (ANR-17-RAR3-0001-01), and l’Institut de Génétique Moléculaire de Montpellier. The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Harrach, B.; Benkő, M. Adenoviridae. In Encyclopedia of Virology; Reference Module in Biomedical Sciences; Human and Animal Viruses; Elsevier: Amsterdam, The Netherlands, 2020; Chapter 00057. [Google Scholar]
  2. Harrach, B.; Tarján, Z.L.; Benkő, M. Adenoviruses across the animal kingdom: A walk in the zoo. FEBS Lett. 2019, 593, 3660–3673. [Google Scholar] [CrossRef]
  3. Lee, C.S.; Bishop, E.S.; Zhang, R.; Yu, X.; Farina, E.M.; Yan, S.; Zhao, C.; Zeng, Z.; Shu, Y.; Wu, X.; et al. Adenovirus-mediated gene delivery: Potential applications for gene and cell-based therapies in the new era of personalized medicine. Genes Dis. 2017, 4, 43–63. [Google Scholar] [CrossRef] [PubMed]
  4. Lion, T. Adenovirus Infections in Immunocompetent and Immunocompromised Patients. Clin. Microbiol. Rev. 2014, 27, 441–462. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Kosulin, K.; Berkowitsch, B.; Matthes, S.; Pichler, H.; Lawitschka, A.; Pötschger, U.; Fritsch, G.; Lion, T. Intestinal Adenovirus Shedding Before Allogeneic Stem Cell Transplantation Is a Risk Factor for Invasive Infection Post-transplant. EBioMedicine 2018, 28, 114–119. [Google Scholar] [CrossRef] [Green Version]
  6. Collins, N.D.; Adhikari, A.; Yang, Y.; Kuschner, R.A.; Karasavvas, N.; Binn, L.N.; Walls, S.D.; Graf, P.C.F.; Myers, C.A.; Jarman, R.G.; et al. Live Oral Adenovirus Type 4 and Type 7 Vaccine Induces Durable Antibody Response. Vaccines 2020, 8, 411. [Google Scholar] [CrossRef] [PubMed]
  7. Rampling, T.; Ewer, K.J.; Bowyer, G.; Bliss, C.M.; Edwards, N.J.; Wright, D.; Payne, R.O.; Venkatraman, N.; de Barra, E.; Snudden, C.M.; et al. Safety and High Level Efficacy of the Combination Malaria Vaccine Regimen of RTS,S/AS01B With Chimpanzee Adenovirus 63 and Modified Vaccinia Ankara Vectored Vaccines Expressing ME-TRAP. J. Infect. Dis. 2016, 214, 772–781. [Google Scholar] [CrossRef] [Green Version]
  8. Sheets, R.L.; Stein, J.; Bailer, R.T.; Koup, R.A.; Andrews, C.; Nason, M.; He, B.; Koo, E.; Trotter, H.; Duffy, C.; et al. Biodistribution and Toxicological Safety of Adenovirus Type 5 and Type 35 Vectored Vaccines Against Human Immunodeficiency Virus-1 (HIV-1), Ebola, or Marburg Are Similar Despite Differing Adenovirus Serotype Vector, Manufacturer’s Construct, or Gene Inserts. J. Immunotoxicol. 2008, 5, 315–335. [Google Scholar] [CrossRef] [PubMed]
  9. Folegatti, P.M.; Ewer, K.J.; Aley, P.K.; Angus, B.; Becker, S.; Belij-Rammerstorfer, S.; Bellamy, D.; Bibi, S.; Bittaye, M.; Clutterbuck, E.A.; et al. Safety and immunogenicity of the ChAdOx1 nCoV-19 vaccine against SARS-CoV-2: A preliminary report of a phase 1/2, single-blind, randomised controlled trial. Lancet 2020, 396, 467–478. [Google Scholar] [CrossRef]
  10. Mercado, N.B.; Zahn, R.; Wegmann, F.; Loos, C.; Chandrashekar, A.; Yu, J.; Liu, J.; Peter, L.; McMahan, K.; Tostanoski, L.H.; et al. Single-shot Ad26 vaccine protects against SARS-CoV-2 in rhesus macaques. Nature 2020, 586, 583–588. [Google Scholar] [CrossRef]
  11. Matz, K.M.; Marzi, A.; Feldmann, H. Ebola vaccine trials: Progress in vaccine safety and immunogenicity. Expert Rev. Vaccines 2019, 18, 1229–1242. [Google Scholar] [CrossRef]
  12. Kremer, E.J. Pros and Cons of Adenovirus-Based SARS-CoV-2 Vaccines. Mol. Ther. 2020, 28, 1–2. [Google Scholar] [CrossRef]
  13. Perreau, M.; Pantaleo, G.; Kremer, E.J. Activation of a dendritic cell–T cell axis by Ad5 immune complexes creates an improved environment for replication of HIV in T cells. J. Exp. Med. 2008, 205, 2717–2725. [Google Scholar] [CrossRef]
  14. Barlan, A.U.; Griffin, T.M.; McGuire, K.A.; Wiethoff, C.M. Adenovirus membrane penetration activates the NLRP3 inflammasome. J. Virol. 2011, 85, 146–155. [Google Scholar] [CrossRef] [Green Version]
  15. Alonso-Padilla, J.; Papp, T.; Kaján, G.L.; Benkő, M.; Havenga, M.; Lemckert, A.; Harrach, B.; Baker, A.H. Development of Novel Adenoviral Vectors to Overcome Challenges Observed With HAdV-5–based Constructs. Mol. Ther. 2016, 24, 6–16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Gaso-Sokac, D.; Kovac, S.; Clifton, J.; Josic, D. Therapeutic plasma proteins—Application of proteomics in process optimization, validation, and analysis of the final product. Electrophoresis 2011, 32, 1104–1117. [Google Scholar] [CrossRef] [PubMed]
  17. Larsson, A.; Hansson, L.-O. Plasma Protein Fractions in Healthy Blood Donors Quantitated by an Automated Multicapillary Electrophoresis System. J. Chromatogr. Sci. 2006, 44, 479–483. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Haney, E.F.; Mansour, S.C.; Hancock, R.E.W. Antimicrobial Peptides: An Introduction. In Antimicrobial Peptides; Methods in Molecular Biology; Hansen, P.R., Ed.; Springer: New York, NY, USA, 2017; Volume 1548, pp. 3–22. ISBN 978-1-4939-6735-3. [Google Scholar]
  19. Schuurhuis, D.H.; van Montfoort, N.; Ioan-Facsinay, A.; Jiawan, R.; Camps, M.; Nouta, J.; Melief, C.J.M.; Verbeek, J.S.; Ossendorp, F. Immune Complex-Loaded Dendritic Cells Are Superior to Soluble Immune Complexes as Antitumor Vaccine. J. Immunol. 2006, 176, 4573–4580. [Google Scholar] [CrossRef] [PubMed]
  20. Mistchenko, A.S.; Diez, R.A.; Mariani, A.L.; Robaldo, J.; Maffey, A.F.; Bayley-Bustamante, G.; Grinstein, S. Cytokines in adenoviral disease in children: Association of interleukin-6, interleukin-8, and tumor necrosis factor alpha levels with clinical outcome. J. Pediatr. 1994, 124, 714–720. [Google Scholar] [CrossRef]
  21. Mistchenko, A.S.; Lenzi, H.L.; Thompson, F.M.; Mota, E.M.; Vidaurreta, S.; Navari, C.; Grinstein, S. Participation of immune complexes in adenovirus infection. Acta Paediatr. 1992, 81, 983–988. [Google Scholar] [CrossRef] [PubMed]
  22. Katzelnick, L.C.; Narvaez, C.; Arguello, S.; Mercado, B.L.; Collado, D.; Ampie, O.; Elizondo, D.; Miranda, T.; Carillo, F.B.; Mercado, J.C.; et al. Zika virus infection enhances future risk of severe dengue disease. Science 2020, 369, 1123–1128. [Google Scholar] [CrossRef] [PubMed]
  23. Katzelnick, L.C.; Bos, S.; Harris, E. Protective and enhancing interactions among dengue viruses 1-4 and Zika virus. Curr. Opin. Virol. 2020, 43, 59–70. [Google Scholar] [CrossRef] [PubMed]
  24. Eichholz, K.; Bru, T.; Tran, T.T.P.; Fernandes, P.; Welles, H.; Mennechet, F.J.D.; Manel, N.; Alves, P.; Perreau, M.; Kremer, E.J. Immune-Complexed Adenovirus Induce AIM2-Mediated Pyroptosis in Human Dendritic Cells. PLoS Pathog 2016, 12, e1005871. [Google Scholar] [CrossRef] [PubMed]
  25. Zaiss, A.K.; Vilaysane, A.; Cotter, M.J.; Clark, S.A.; Meijndert, H.C.; Colarusso, P.; Yates, R.M.; Petrilli, V.; Tschopp, J.; Muruve, D.A. Antiviral Antibodies Target Adenovirus to Phagolysosomes and Amplify the Innate Immune Response. J. Immunol. 2009, 182, 7058–7068. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Somanathan, S.; Calcedo, R.; Wilson, J.M. Adenovirus-Antibody Complexes Contributed to Lethal Systemic Inflammation in a Gene Therapy Trial. Mol. Ther. 2020, 28, 784–793. [Google Scholar] [CrossRef] [PubMed]
  27. Tran, T.T.P.; Eichholz, K.; Amelio, P.; Moyer, C.; Nemerow, G.R.; Perreau, M.; Mennechet, F.J.D.; Kremer, E.J. Humoral immune response to adenovirus induce tolerogenic bystander dendritic cells that promote generation of regulatory T cells. PLoS Pathog 2018, 14, e1007127. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Korokhov, N.; de Gruijl, T.D.; Aldrich, W.A.; Triozzi, P.L.; Banerjee, P.T.; Gillies, S.D.; Curiel, T.J.; Douglas, J.T.; Scheper, R.J.; Curiel, D.T. High efficiency transduction of dendritic cells by adenoviral vectors targeted to DC-SIGN. Cancer Biol. Ther. 2005, 4, 289–294. [Google Scholar] [CrossRef] [Green Version]
  29. Segura, E.; Amigorena, S. Cross-Presentation in Mouse and Human Dendritic Cells. In Advances in Immunology; Elsevier: Amsterdam, The Netherlands, 2015; Volume 127, pp. 1–31. ISBN 978-0-12-802245-0. [Google Scholar]
  30. Mallery, D.L.; McEwan, W.A.; Bidgood, S.R.; Towers, G.J.; Johnson, C.M.; James, L.C. Antibodies mediate intracellular immunity through tripartite motif-containing 21 (TRIM21). Proc. Natl. Acad. Sci. USA 2010, 107, 19985–19990. [Google Scholar] [CrossRef] [Green Version]
  31. Fletcher, A.J.; James, L.C. Coordinated Neutralization and Immune Activation by the Cytosolic Antibody Receptor TRIM21. J. Virol. 2016, 90, 4856–4859. [Google Scholar] [CrossRef] [Green Version]
  32. Bottermann, M.; Foss, S.; van Tienen, L.M.; Vaysburd, M.; Cruickshank, J.; O’Connell, K.; Clark, J.; Mayes, K.; Higginson, K.; Hirst, J.C.; et al. TRIM21 mediates antibody inhibition of adenovirus-based gene delivery and vaccination. Proc. Natl. Acad. Sci. USA 2018, 115, 10440–10445. [Google Scholar] [CrossRef] [Green Version]
  33. McEwan, W.A.; Hauler, F.; Williams, C.R.; Bidgood, S.R.; Mallery, D.L.; Crowther, R.A.; James, L.C. Regulation of Virus Neutralization and the Persistent Fraction by TRIM21. J. Virol. 2012, 86, 8482–8491. [Google Scholar] [CrossRef] [Green Version]
  34. McElrath, M.J.; De Rosa, S.C.; Moodie, Z.; Dubey, S.; Kierstead, L.; Janes, H.; Defawe, O.D.; Carter, D.K.; Hural, J.; Akondy, R.; et al. HIV-1 vaccine-induced immunity in the test-of-concept Step Study: A case–cohort analysis. Lancet 2008, 372, 1894–1905. [Google Scholar] [CrossRef] [Green Version]
  35. Buchbinder, S.P.; Mehrotra, D.V.; Duerr, A.; Fitzgerald, D.W.; Mogg, R.; Li, D.; Gilbert, P.B.; Lama, J.R.; Marmor, M.; del Rio, C.; et al. Efficacy assessment of a cell-mediated immunity HIV-1 vaccine (the Step Study): A double-blind, randomised, placebo-controlled, test-of-concept trial. Lancet 2008, 372, 1881–1893. [Google Scholar] [CrossRef] [Green Version]
  36. Calabro, S.; Tortoli, M.; Baudner, B.C.; Pacitto, A.; Cortese, M.; O’Hagan, D.T.; De Gregorio, E.; Seubert, A.; Wack, A. Vaccine adjuvants alum and MF59 induce rapid recruitment of neutrophils and monocytes that participate in antigen transport to draining lymph nodes. Vaccine 2011, 29, 1812–1823. [Google Scholar] [CrossRef] [PubMed]
  37. Benlahrech, A.; Harris, J.; Meiser, A.; Papagatsias, T.; Hornig, J.; Hayes, P.; Lieber, A.; Athanasopoulos, T.; Bachy, V.; Csomor, E.; et al. Adenovirus vector vaccination induces expansion of memory CD4 T cells with a mucosal homing phenotype that are readily susceptible to HIV-1. Proc. Natl. Acad. Sci. USA 2009, 106, 19940–19945. [Google Scholar] [CrossRef] [Green Version]
  38. Baker, A.H.; Herzog, R.W. Did Dendritic Cell Activation, Induced by Adenovirus-Antibody Complexes, Play a Role in the Death of Jesse Gelsinger? Mol. Ther. 2020, 28, 704–706. [Google Scholar] [CrossRef]
  39. Zhuang, Q.-C.; Sun, C.-J.; Li, F.; Feng, L.-Q.; Liu, Y.-C.; Chen, L. Monocytes are more susceptible in vitro in rhesus macaques of Chinese origin to recombinant serotype 5 adenovirus with higher preexisting vector specific neutralizing antibody titer. Bing Du Xue Bao 2008, 24, 383–389. [Google Scholar]
  40. Ricklin, D.; Hajishengallis, G.; Yang, K.; Lambris, J.D. Complement: A key system for immune surveillance and homeostasis. Nat. Immunol. 2010, 11, 785–797. [Google Scholar] [CrossRef] [Green Version]
  41. Ochsenbein, A.F.; Zinkernagel, R.M. Natural antibodies and complement link innate and acquired immunity. Immunol. Today 2000, 21, 624–630. [Google Scholar] [CrossRef]
  42. Huber-Lang, M.; Sarma, J.V.; Zetoune, F.S.; Rittirsch, D.; Neff, T.A.; McGuire, S.R.; Lambris, J.D.; Warner, R.L.; Flierl, M.A.; Hoesel, L.M.; et al. Generation of C5a in the absence of C3: A new complement activation pathway. Nat. Med. 2006, 12, 682–687. [Google Scholar] [CrossRef]
  43. Gasque, P. Complement: A unique innate immune sensor for danger signals. Mol. Immunol. 2004, 41, 1089–1098. [Google Scholar] [CrossRef]
  44. Guo, R.-F.; Ward, P.A. Role of C5A in Inflammatory Responses. Annu. Rev. Immunol. 2005, 23, 821–852. [Google Scholar] [CrossRef] [PubMed]
  45. Cichon, G.; Boeckh-Herwig, S.; Schmidt, H.; Wehnes, E.; Müller, T.; Pring-Akerblom, P.; Burger, R. Complement activation by recombinant adenoviruses. Gene 2001, 8, 1794–1800. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Jiang, H.; Wang, Z.; Serra, D.; Frank, M.M.; 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, 1140–1142. [Google Scholar] [CrossRef] [PubMed]
  47. Bottermann, M.; Foss, S.; Caddy, S.L.; Clift, D.; van Tienen, L.M.; Vaysburd, M.; Cruickshank, J.; O’Connell, K.; Clark, J.; Mayes, K.; et al. Complement C4 Prevents Viral Infection through Capsid Inactivation. Cell Host Microbe 2019, 25, 617–629. [Google Scholar] [CrossRef] [Green Version]
  48. Perreau, M.; Guérin, M.-C.; Drouet, C.; Kremer, E.J. Interactions Between Human Plasma Components and A Xenogenic Adenovirus Vector: Reduced Immunogenicity During Gene Transfer. Mol. Ther. 2007, 15, 1998–2007. [Google Scholar] [CrossRef]
  49. Verschoor, A.; Karsten, C.M.; Broadley, S.P.; Laumonnier, Y.; Köhl, J. Old dogs-new tricks: Immunoregulatory properties of C3 and C5 cleavage fragments. Immunol. Rev. 2016, 274, 112–126. [Google Scholar] [CrossRef]
  50. Czermak, B.J.; Sarma, V.; Bless, N.M.; Schmal, H.; Friedl, H.P.; Ward, P.A. In vitro and in vivo dependency of chemokine generation on C5a and TNF-alpha. J. Immunol. 1999, 162, 2321–2325. [Google Scholar]
  51. Frank, M.M.; Fries, L.F. The role of complement in inflammation and phagocytosis. Immunol. Today 1991, 12, 322–326. [Google Scholar] [CrossRef]
  52. Cotter, M.J.; Zaiss, A.K.; Muruve, D.A. Neutrophils Interact with Adenovirus Vectors via Fc Receptors and Complement Receptor 1. J. Virol. 2005, 79, 14622–14631. [Google Scholar] [CrossRef] [Green Version]
  53. Soruri, A.; Kiafard, Z.; Dettmer, C.; Riggert, J.; Köhl, J.; Zwirner, J. IL-4 Down-Regulates Anaphylatoxin Receptors in Monocytes and Dendritic Cells and Impairs Anaphylatoxin-Induced Migration In Vivo. J. Immunol. 2003, 170, 3306–3314. [Google Scholar] [CrossRef] [Green Version]
  54. Zhou, W. The new face of anaphylatoxins in immune regulation. Immunobiology 2012, 217, 225–234. [Google Scholar] [CrossRef] [PubMed]
  55. Asgari, E.; Le Friec, G.; Yamamoto, H.; Perucha, E.; Sacks, S.S.; Köhl, J.; Cook, H.T.; Kemper, C. C3a modulates IL-1β secretion in human monocytes by regulating ATP efflux and subsequent NLRP3 inflammasome activation. Blood 2013, 122, 3473–3481. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Suresh, R.; Chandrasekaran, P.; Sutterwala, F.S.; Mosser, D.M. Complement-mediated “bystander” damage initiates host NLRP3 inflammasome activation. J. Cell Sci. 2016, 129, 1928–1939. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Hawlisch, H.; Belkaid, Y.; Baelder, R.; Hildeman, D.; Gerard, C.; Köhl, J. C5a Negatively Regulates Toll-like Receptor 4-Induced Immune Responses. Immunity 2005, 22, 415–426. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Karsten, C.M.; Pandey, M.K.; Figge, J.; Kilchenstein, R.; Taylor, P.R.; Rosas, M.; McDonald, J.U.; Orr, S.J.; Berger, M.; Petzold, D.; et al. Anti-inflammatory activity of IgG1 mediated by Fc galactosylation and association of FcγRIIB and dectin-1. Nat. Med. 2012, 18, 1401–1406. [Google Scholar] [CrossRef]
  59. Kohl, J. A regulatory role for the C5a anaphylatoxin in type 2 immunity in asthma. J. Clin. Investig. 2006, 116, 783–796. [Google Scholar] [CrossRef] [Green Version]
  60. Strainic, M.G.; Shevach, E.M.; An, F.; Lin, F.; Medof, M.E. Absence of signaling into CD4+ cells via C3aR and C5aR enables autoinductive TGF-β1 signaling and induction of Foxp3+ regulatory T cells. Nat. Immunol. 2013, 14, 162–171. [Google Scholar] [CrossRef] [Green Version]
  61. Verschoor, A.; Neuenhahn, M.; Navarini, A.A.; Graef, P.; Plaumann, A.; Seidlmeier, A.; Nieswandt, B.; Massberg, S.; Zinkernagel, R.M.; Hengartner, H.; et al. A platelet-mediated system for shuttling blood-borne bacteria to CD8α+ dendritic cells depends on glycoprotein GPIb and complement C3. Nat. Immunol. 2011, 12, 1194–1201. [Google Scholar] [CrossRef]
  62. Seiradake, E.; Henaff, D.; Wodrich, H.; Billet, O.; Perreau, M.; Hippert, C.; Mennechet, F.; Schoehn, G.; Lortat-Jacob, H.; Dreja, H.; et al. The Cell Adhesion Molecule “CAR” and Sialic Acid on Human Erythrocytes Influence Adenovirus In Vivo Biodistribution. PLoS Pathog 2009, 5, e1000277. [Google Scholar] [CrossRef] [Green Version]
  63. Hess, C.; Schifferli, J.A. Immune Adherence Revisited: Novel Players in an Old Game. Physiology 2003, 18, 104–108. [Google Scholar] [CrossRef] [Green Version]
  64. Paccaud, J.P.; Carpentier, J.L.; Schifferli, J.A. Direct evidence for the clustered nature of complement receptors type 1 on the erythrocyte membrane. J. Immunol. 1988, 141, 3889–3894. [Google Scholar] [PubMed]
  65. Ward, P.A. Sepsis, apoptosis and complement. Biochem. Pharm. 2008, 76, 1383–1388. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Guaní-Guerra, E.; Santos-Mendoza, T.; Lugo-Reyes, S.O.; Terán, L.M. Antimicrobial peptides: General overview and clinical implications in human health and disease. Clin. Immunol. 2010, 135, 1–11. [Google Scholar] [CrossRef]
  67. Lai, Y.; Gallo, R.L. AMPed up immunity: How antimicrobial peptides have multiple roles in immune defense. Trends Immunol. 2009, 30, 131–141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Nguyen, L.T.; Haney, E.F.; Vogel, H.J. The expanding scope of antimicrobial peptide structures and their modes of action. Trends Biotechnol. 2011, 29, 464–472. [Google Scholar] [CrossRef] [PubMed]
  69. Da Cunha, N.B.; Cobacho, N.B.; Viana, J.F.C.; Lima, L.A.; Sampaio, K.B.O.; Dohms, S.S.M.; Ferreira, A.C.R.; de la Fuente-Núñez, C.; Costa, F.F.; Franco, O.L.; et al. The next generation of antimicrobial peptides (AMPs) as molecular therapeutic tools for the treatment of diseases with social and economic impacts. Drug Discov. Today 2017, 22, 234–248. [Google Scholar] [CrossRef]
  70. Dorschner, R.A.; Pestonjamasp, V.K.; Tamakuwala, S.; Ohtake, T.; Rudisill, J.; Nizet, V.; Agerberth, B.; Gudmundsson, G.H.; Gallo, R.L. Cutaneous Injury Induces the Release of Cathelicidin Anti-Microbial Peptides Active Against Group A Streptococcus. J. Investig. Dermatol. 2001, 117, 91–97. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Kundhavai Natchiar, S.; Venkataraman, S.; Mullen, T.-M.; Nemerow, G.R.; Reddy, V.S. Revised Crystal Structure of Human Adenovirus Reveals the Limits on Protein IX Quasi-Equivalence and on Analyzing Large Macromolecular Complexes. J. Mol. Biol. 2018, 430, 4132–4141. [Google Scholar] [CrossRef]
  72. Nemerow, G.R.; Stewart, P.L.; Reddy, V.S. Structure of human adenovirus. Curr. Opin. Virol. 2012, 2, 115–121. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Reddy, V.S.; Natchiar, S.K.; Stewart, P.L.; Nemerow, G.R. Crystal Structure of Human Adenovirus at 3.5 A Resolution. Science 2010, 329, 1071–1075. [Google Scholar] [CrossRef] [Green Version]
  74. Alba, R.; Bradshaw, A.C.; Parker, A.L.; Bhella, D.; Waddington, S.N.; Nicklin, S.A.; van Rooijen, N.; Custers, J.; Goudsmit, J.; Barouch, D.H.; et al. Identification of coagulation factor (F)X binding sites on the adenovirus serotype 5 hexon: Effect of mutagenesis on FX interactions and gene transfer. Blood 2009, 114, 965–971. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Persson, B.D.; Lenman, A.; Frängsmyr, L.; Schmid, M.; Ahlm, C.; Plückthun, A.; Jenssen, H.; Arnberg, N. Lactoferrin-Hexon Interactions Mediate CAR-Independent Adenovirus Infection of Human Respiratory Cells. J. Virol. 2020, 94, e00542-20. [Google Scholar] [CrossRef] [PubMed]
  76. Presicce, P.; Giannelli, S.; Taddeo, A.; Villa, M.L.; Della Bella, S. Human defensins activate monocyte-derived dendritic cells, promote the production of proinflammatory cytokines, and up-regulate the surface expression of CD91. J. Leukoc. Biol. 2009, 86, 941–948. [Google Scholar] [CrossRef] [PubMed]
  77. Territo, M.C.; Ganz, T.; Selsted, M.E.; Lehrer, R. Monocyte-chemotactic activity of defensins from human neutrophils. J. Clin. Investig. 1989, 84, 2017–2020. [Google Scholar] [CrossRef]
  78. Chertov, O.; Michiel, D.F.; Xu, L.; Wang, J.M.; Tani, K.; Murphy, W.J.; Longo, D.L.; Taub, D.D.; Oppenheim, J.J. Identification of Defensin-1, Defensin-2, and CAP37/Azurocidin as T-cell Chemoattractant Proteins Released from Interleukin-8-stimulated Neutrophils. J. Biol. Chem. 1996, 271, 2935–2940. [Google Scholar] [CrossRef] [Green Version]
  79. Yang, D.; Chen, Q.; Chertov, O.; Oppenheim, J.J. Human neutrophil defensins selectively chemoattract naive T and immature dendritic cells. J. Leukoc. Biol. 2000, 68, 9–14. [Google Scholar]
  80. Yang, D.; Chertov, O.; Bykovskaia, S.N.; Chen, Q.; Buffo, M.J.; Shogan, J.; Anderson, M.; Schröder, J.M.; Wang, J.M.; Howard, O.M.Z.; et al. β-Defensins: Linking Innate and Adaptive Immunity Through Dendritic and T Cell CCR6. Science 1999, 286, 525–528. [Google Scholar] [CrossRef]
  81. Niyonsaba, F.; Iwabuchi, K.; Matsuda, H.; Ogawa, H.; Nagaoka, I. Epithelial cell-derived human β-defensin-2 acts as a chemotaxin for mast cells through a pertussis toxin-sensitive and phospholipase C-dependent pathway. Int. Immunol. 2002, 14, 421–426. [Google Scholar] [CrossRef] [Green Version]
  82. García, J.-R.C.; Krause, A.; Schulz, S.; Rodríguez-Jiménez, F.-J.; Klüver, E.; Adermann, K.; Forssmann, U.; Frimpong-Boateng, A.; Bals, R.; Forssmann, W.-G. Human β-defensin 4: A novel inducible peptide with a specific salt-sensitive spectrum of antimicrobial activity. FASEB J. 2001, 15, 1819–1821. [Google Scholar] [CrossRef] [Green Version]
  83. Funderburg, N.; Lederman, M.M.; Feng, Z.; Drage, M.G.; Jadlowsky, J.; Harding, C.V.; Weinberg, A.; Sieg, S.F. Human -defensin-3 activates professional antigen-presenting cells via Toll-like receptors 1 and 2. Proc. Natl. Acad. Sci. USA 2007, 104, 18631–18635. [Google Scholar] [CrossRef] [Green Version]
  84. Spadaro, M.; Caorsi, C.; Ceruti, P.; Varadhachary, A.; Forni, G.; Pericle, F.; Giovarelli, M. Lactoferrin, a major defense protein of innate immunity, is a novel maturation factor for human dendritic cells. FASEB J. 2016, 22, 2747–2757. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. De la Rosa, G.; Yang, D.; Tewary, P.; Varadhachary, A.; Oppenheim, J.J. Lactoferrin Acts as an Alarmin to Promote the Recruitment and Activation of APCs and Antigen-Specific Immune Responses. J. Immunol. 2008, 180, 6868–6876. [Google Scholar] [CrossRef] [PubMed]
  86. Adams, W.C.; Bond, E.; Havenga, M.J.E.; Holterman, L.; Goudsmit, J.; Karlsson Hedestam, G.B.; Koup, R.A.; Loré, K.; Hedestam, G.B.K.; Koup, R.A.; et al. Adenovirus serotype 5 infects human dendritic cells via a coxsackievirus-adenovirus receptor-independent receptor pathway mediated by lactoferrin and DC-SIGN. J. Gen. Virol. 2009, 90, 1600–1610. [Google Scholar] [CrossRef]
  87. Spadaro, M.; Montone, M.; Arigoni, M.; Cantarella, D.; Forni, G.; Pericle, F.; Pascolo, S.; Calogero, R.A.; Cavallo, F. Recombinant human lactoferrin induces human and mouse dendritic cell maturation via Toll-like receptors 2 and 4. FASEB J. 2014, 28, 416–429. [Google Scholar] [CrossRef]
  88. Van der Does, A.M.; Hensbergen, P.J.; Bogaards, S.J.; Cansoy, M.; Deelder, A.M.; van Leeuwen, H.C.; Drijfhout, J.W.; van Dissel, J.T.; Nibbering, P.H. The Human Lactoferrin-Derived Peptide hLF1-11 Exerts Immunomodulatory Effects by Specific Inhibition of Myeloperoxidase Activity. J. Immunol. 2012, 188, 5012–5019. [Google Scholar] [CrossRef] [Green Version]
  89. Yang, D.; Chen, Q.; Schmidt, A.P.; Anderson, G.M.; Wang, J.M.; Wooters, J.; Oppenheim, J.J.; Chertov, O. Ll-37, the Neutrophil Granule–And Epithelial Cell–Derived Cathelicidin, Utilizes Formyl Peptide Receptor–Like 1 (Fprl1) as a Receptor to Chemoattract Human Peripheral Blood Neutrophils, Monocytes, and T Cells. J. Exp. Med. 2000, 192, 1069–1074. [Google Scholar] [CrossRef]
  90. Zhang, X.; Bajic, G.; Andersen, G.R.; Christiansen, S.H.; Vorup-Jensen, T. The cationic peptide LL-37 binds Mac-1 (CD11b/CD18) with a low dissociation rate and promotes phagocytosis. Biochim. Biophys. Acta Proteins Proteom. 2016, 1864, 471–478. [Google Scholar] [CrossRef]
  91. Kandler, K.; Shaykhiev, R.; Kleemann, P.; Klescz, F.; Lohoff, M.; Vogelmeier, C.; Bals, R. The anti-microbial peptide LL-37 inhibits the activation of dendritic cells by TLR ligands. Int. Immunol. 2006, 18, 1729–1736. [Google Scholar] [CrossRef] [Green Version]
  92. Hurtado, P.; Peh, C.A. LL-37 promotes rapid sensing of CpG oligodeoxynucleotides by B lymphocytes and plasmacytoid dendritic cells. J. Immunol. 2010, 184, 1425–1435. [Google Scholar] [CrossRef] [Green Version]
  93. Chamilos, G.; Gregorio, J.; Meller, S.; Lande, R.; Kontoyiannis, D.P.; Modlin, R.L.; Gilliet, M. Cytosolic sensing of extracellular self-DNA transported into monocytes by the antimicrobial peptide LL37. Blood 2012, 120, 3699–3707. [Google Scholar] [CrossRef] [Green Version]
  94. Davidson, D.J.; Currie, A.J.; Reid, G.S.D.; Bowdish, D.M.E.; MacDonald, K.L.; Ma, R.C.; Hancock, R.E.W.; Speert, D.P. The cationic antimicrobial peptide LL-37 modulates dendritic cell differentiation and dendritic cell-induced T cell polarization. J. Immunol. 2004, 172, 1146–1156. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Smith, J.G.; Nemerow, G.R. Mechanism of adenovirus neutralization by Human alpha-defensins. Cell Host Microbe 2008, 3, 11–19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Flatt, J.W.; Kim, R.; Smith, J.G.; Nemerow, G.R.; Stewart, P.L. An intrinsically disordered region of the adenovirus capsid is implicated in neutralization by human alpha defensin 5. PLoS ONE 2013, 8, e61571. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Nguyen, E.K.; Nemerow, G.R.; Smith, J.G. Direct evidence from single-cell analysis that human α-defensins block adenovirus uncoating to neutralize infection. J. Virol. 2010, 84, 4041–4049. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Smith, J.G.; Silvestry, M.; Lindert, S.; Lu, W.; Nemerow, G.R.; Stewart, P.L. Insight into the mechanisms of adenovirus capsid disassembly from studies of defensin neutralization. PLoS Pathog. 2010, 6, e1000959. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Skrygan, M.; Schäfer, H.; Schmidt, C.; Schmidt, W.E.; Bastian, A. Inhibition of adenovirus infection by the bronchoalveolar lavage supernatant in vitro. Eur. J. Med. Res. 2003, 8, 519–524. [Google Scholar]
  100. Tartaglia, L.J.; Badamchi-Zadeh, A.; Abbink, P.; Blass, E.; Aid, M.; Gebre, M.S.; Li, Z.; Pastores, K.C.; Trott, S.; Gupte, S.; et al. Alpha-defensin 5 differentially modulates adenovirus vaccine vectors from different serotypes in vivo. PLoS Pathog. 2019, 15, e1008180. [Google Scholar] [CrossRef]
  101. Priyadharshini, V.S.; Ramírez-Jiménez, F.; Molina-Macip, M.; Renteria-Rosales, C.; Santiago-Cruz, J.; Zarate-Segura, P.; Lara-Padilla, E.; Teran, L.M. Human Neutrophil Defensin-1, -3, and -4 Are Elevated in Nasal Aspirates from Children with Naturally Occurring Adenovirus Infection. Can. Respir. J. 2018, 2018, 1–6. [Google Scholar] [CrossRef] [Green Version]
  102. Vragniau, C.; Hübner, J.-M.; Beidler, P.; Gil, S.; Saydaminova, K.; Lu, Z.-Z.; Yumul, R.; Wang, H.; Richter, M.; Sova, P.; et al. Studies on the Interaction of Tumor-Derived HD5 Alpha Defensins with Adenoviruses and Implications for Oncolytic Adenovirus Therapy. J. Virol. 2017, 91, e02030-16. [Google Scholar] [CrossRef] [Green Version]
  103. Johansson, C.; Jonsson, M.; Marttila, M.; Persson, D.; Fan, X.L.; Skog, J.; Frangsmyr, L.; Wadell, G.; Arnberg, N. Adenoviruses use lactoferrin as a bridge for CAR-independent binding to and infection of epithelial cells. J. Virol. 2007, 81, 954–963. [Google Scholar] [CrossRef] [Green Version]
  104. Loustalot, F.; Kremer, E.J.; Salinas, S. Membrane Dynamics and Signaling of the Coxsackievirus and Adenovirus Receptor. In International Review of Cell and Molecular Biology; Elsevier: Amsterdam, The Netherlands, 2016; Volume 322, pp. 331–362. ISBN 978-0-12-804809-2. [Google Scholar]
  105. Arnberg, N. Adenovirus receptors: Implications for targeting of viral vectors. Trends Pharmacol. Sci. 2012, 33, 442–448. [Google Scholar] [CrossRef]
  106. Günther, P.S.; Mikeler, E.; Hamprecht, K.; Schneider-Schaulies, J.; Jahn, G.; Dennehy, K.M. CD209/DC-SIGN mediates efficient infection of monocyte-derived dendritic cells by clinical adenovirus 2C isolates in the presence of bovine lactoferrin. J. Gen. Virol. 2011, 92, 1754–1759. [Google Scholar] [CrossRef] [PubMed]
  107. Uchio, E.; Inoue, H.; Kadonosono, K. Anti-adenoviral effects of human cationic antimicrobial protein-18/LL-37, an antimicrobial peptide, by quantitative polymerase chain reaction. Korean J. Ophthalmol. 2013, 27, 199–203. [Google Scholar] [CrossRef] [PubMed]
  108. Gordon, Y.J.; Huang, L.C.; Romanowski, E.G.; Yates, K.A.; Proske, R.J.; McDermott, A.M. Human cathelicidin (LL-37), a multifunctional peptide, is expressed by ocular surface epithelia and has potent antibacterial and antiviral activity. Curr. Eye Res. 2005, 30, 385–394. [Google Scholar] [CrossRef] [PubMed]
  109. Palta, S.; Saroa, R.; Palta, A. Overview of the coagulation system. Indian J. Anaesth. 2014, 58, 515. [Google Scholar] [CrossRef] [PubMed]
  110. Kalyuzhniy, O.; Di Paolo, N.C.; Silvestry, M.; Hofherr, S.E.; Barry, M.A.; Stewart, P.L.; Shayakhmetov, D.M. Adenovirus serotype 5 hexon is critical for virus infection of hepatocytes in vivo. Proc. Natl. Acad. Sci. USA 2008, 105, 5483–5488. [Google Scholar] [CrossRef] [Green Version]
  111. Waddington, S.N.; McVey, J.H.; Bhella, D.; Parker, A.L.; Barker, K.; Atoda, H.; Pink, R.; Buckley, S.M.K.; Greig, J.A.; Denby, L.; et al. Adenovirus Serotype 5 Hexon Mediates Liver Gene Transfer. Cell 2008, 132, 397–409. [Google Scholar] [CrossRef] [Green Version]
  112. Parker, A.L.; Waddington, S.N.; Nicol, C.G.; Shayakhmetov, D.M.; Buckley, S.M.; Denby, L.; Kemball-Cook, G.; Ni, S.; Lieber, A.; McVey, J.H.; et al. Multiple vitamin K-dependent coagulation zymogens promote adenovirus-mediated gene delivery to hepatocytes. Blood 2006, 108, 2554–2561. [Google Scholar] [CrossRef] [Green Version]
  113. Lopez-Gordo, E.; Denby, L.; Nicklin, S.A.; Baker, A.H. The importance of coagulation factors binding to adenovirus: Historical perspectives and implications for gene delivery. Expert Opin. Drug Deliv. 2014, 11, 1795–1813. [Google Scholar] [CrossRef]
  114. Allen, R.J.; Byrnes, A.P. Interaction of adenovirus with antibodies, complement, and coagulation factors. FEBS Lett. 2019, 593, 3449–3460. [Google Scholar] [CrossRef]
  115. Jonsson, M.I.; Lenman, A.E.; Frängsmyr, L.; Nyberg, C.; Abdullahi, M.; Arnberg, N. Coagulation Factors IX and X Enhance Binding and Infection of Adenovirus Types 5 and 31 in Human Epithelial Cells. J. Virol. 2009, 83, 3816–3825. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Eichholz, K.; Mennechet, F.J.D.; Kremer, E.J. Human Coagulation Factor X-Adenovirus Type 5 Complexes Poorly Stimulate an Innate Immune Response in Human Mononuclear Phagocytes. J. Virol. 2015, 89, 2884–2891. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Alba, R.; Bradshaw, A.C.; Mestre-Francés, N.; Verdier, J.-M.; Henaff, D.; Baker, A.H. Coagulation factor X mediates adenovirus type 5 liver gene transfer in non-human primates (Microcebus murinus). Gene 2012, 19, 109–113. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Hirschowitz, E.; Weaver, J.; Hidalgo, G.; Doherty, D. Murine dendritic cells infected with adenovirus vectors show signs of activation. Gene 2000, 7, 1112–1120. [Google Scholar] [CrossRef] [Green Version]
  119. Doronin, K.; Flatt, J.W.; Di Paolo, N.C.; Khare, R.; Kalyuzhniy, O.; Acchione, M.; Sumida, J.P.; Ohto, U.; Shimizu, T.; Akashi-Takamura, S.; et al. Coagulation Factor X Activates Innate Immunity to Human Species C Adenovirus. Science 2012, 338, 795–798. [Google Scholar] [CrossRef] [Green Version]
  120. Gounder, A.P.; Myers, N.D.; Treuting, P.M.; Bromme, B.A.; Wilson, S.S.; Wiens, M.E.; Lu, W.; Ouellette, A.J.; Spindler, K.R.; Parks, W.C.; et al. Defensins Potentiate a Neutralizing Antibody Response to Enteric Viral Infection. PLoS Pathog. 2016, 12, e1005474. [Google Scholar] [CrossRef] [Green Version]
  121. Kijlstra, A. Modulation of classical C3 convertase of complement by tear lactoferrin. Immunology 1982, 47, 263–270. [Google Scholar]
  122. van den Berg, R.H.; Faber-Krol, M.C.; van Wetering, S.; Hiemstra, P.S.; Daha, M.R. Inhibition of Activation of the Classical Pathway of Complement by Human Neutrophil Defensins. Blood 1998, 92, 3898–3903. [Google Scholar] [CrossRef]
  123. Xu, Z.; Qiu, Q.; Tian, J.; Smith, J.S.; Conenello, G.M.; Morita, T.; Byrnes, A.P. Coagulation factor X shields adenovirus type 5 from attack by natural antibodies and complement. Nat. Med. 2013, 19, 452–457. [Google Scholar] [CrossRef]
  124. Krutzke, L.; Prill, J.M.; Engler, T.; Schmidt, C.Q.; Xu, Z.; Byrnes, A.P.; Simmet, T.; Kreppel, F. Substitution of blood coagulation factor X-binding to Ad5 by position-specific PEGylation: Preventing vector clearance and preserving infectivity. J. Control. Release 2016, 235, 379–392. [Google Scholar] [CrossRef] [Green Version]
  125. Hill, C.; Carlisle, R. Achieving systemic delivery of oncolytic viruses. Expert Opin. Drug Deliv. 2019, 16, 607–620. [Google Scholar] [CrossRef] [Green Version]
  126. Hagedorn, C.; Kreppel, F. Capsid Engineering of Adenovirus Vectors: Overcoming Early Vector–Host Interactions for Therapy. Hum. Gene Ther. 2017, 28, 820–832. [Google Scholar] [CrossRef] [PubMed]
  127. Yoon, A.-R.; Hong, J.; Kim, S.W.; Yun, C.-O. Redirecting adenovirus tropism by genetic, chemical, and mechanical modification of the adenovirus surface for cancer gene therapy. Expert Opin. Drug Deliv. 2016, 13, 843–858. [Google Scholar] [CrossRef]
  128. Cawston, T.E.; Young, D.A. Proteinases involved in matrix turnover during cartilage and bone breakdown. Cell Tissue Res. 2010, 339, 221–235. [Google Scholar] [CrossRef] [PubMed]
  129. Lavilla-Alonso, S.; Bauer, M.M.T.; Abo-Ramadan, U.; Ristimäki, A.; Halavaara, J.; Desmond, R.A.; Wang, D.; Escutenaire, S.; Ahtiainen, L.; Saksela, K.; et al. Macrophage metalloelastase (MME) as adjuvant for intra-tumoral injection of oncolytic adenovirus and its influence on metastases development. Cancer Gene Ther. 2012, 19, 126–134. [Google Scholar] [CrossRef] [PubMed]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chéneau, C.; Kremer, E.J. Adenovirus—Extracellular Protein Interactions and Their Impact on Innate Immune Responses by Human Mononuclear Phagocytes. Viruses 2020, 12, 1351. https://0-doi-org.brum.beds.ac.uk/10.3390/v12121351

AMA Style

Chéneau C, Kremer EJ. Adenovirus—Extracellular Protein Interactions and Their Impact on Innate Immune Responses by Human Mononuclear Phagocytes. Viruses. 2020; 12(12):1351. https://0-doi-org.brum.beds.ac.uk/10.3390/v12121351

Chicago/Turabian Style

Chéneau, Coraline, and Eric J. Kremer. 2020. "Adenovirus—Extracellular Protein Interactions and Their Impact on Innate Immune Responses by Human Mononuclear Phagocytes" Viruses 12, no. 12: 1351. https://0-doi-org.brum.beds.ac.uk/10.3390/v12121351

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop