Next Article in Journal
A mRNA Vaccine for Crimean–Congo Hemorrhagic Fever Virus Expressing Non-Fusion GnGc Using NSm Linker Elicits Unexpected Immune Responses in Mice
Next Article in Special Issue
DMSO and Its Role in Differentiation Impact Efficacy of Human Adenovirus (HAdV) Infection in HepaRG Cells
Previous Article in Journal
Deletion of the EP402R Gene from the Genome of African Swine Fever Vaccine Strain ASFV-G-∆I177L Provides the Potential Capability of Differentiating between Infected and Vaccinated Animals
Previous Article in Special Issue
Evolving Horizons: Adenovirus Vectors’ Timeless Influence on Cancer, Gene Therapy and Vaccines
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advances of Recombinant Adenoviral Vectors in Preclinical and Clinical Applications

by
Luca Scarsella
1,2,3,*,
Eric Ehrke-Schulz
2,
Michael Paulussen
4,
Serge C. Thal
1,
Anja Ehrhardt
2 and
Malik Aydin
2,3,4,5,*
1
Department of Anesthesiology, Center for Clinical and Translational Research, Helios University Hospital Wuppertal, Witten/Herdecke University, 42283 Wuppertal, Germany
2
Virology and Microbiology, Center for Biomedical Education and Research (ZBAF), Department Human Medicine, Faculty of Health, Witten/Herdecke University, 58453 Witten, Germany
3
Laboratory of Experimental Pediatric Pneumology and Allergology, Center for Biomedical Education and Science (ZBAF), Department of Human Medicine, Faculty of Medicine, Witten/Herdecke University, 58453 Witten, Germany
4
Chair of Pediatrics, University Children’s Hospital, Vestische Kinder- und Jugendklinik Datteln, Witten/Herdecke University, 45711 Datteln, Germany
5
Institute for Medical Laboratory Diagnostics, Center for Clinical and Translational Research, Helios University Hospital Wuppertal, Witten/Herdecke University, 42283 Wuppertal, Germany
*
Authors to whom correspondence should be addressed.
Viruses 2024, 16(3), 377; https://doi.org/10.3390/v16030377
Submission received: 31 December 2023 / Revised: 14 February 2024 / Accepted: 19 February 2024 / Published: 28 February 2024
(This article belongs to the Special Issue Research and Clinical Application of Adenovirus (AdV), Volume II)

Abstract

:
Adenoviruses (Ad) have the potential to induce severe infections in vulnerable patient groups. Therefore, understanding Ad biology and antiviral processes is important to comprehend the signaling cascades during an infection and to initiate appropriate diagnostic and therapeutic interventions. In addition, Ad vector-based vaccines have revealed significant potential in generating robust immune protection and recombinant Ad vectors facilitate efficient gene transfer to treat genetic diseases and are used as oncolytic viruses to treat cancer. Continuous improvements in gene delivery capacity, coupled with advancements in production methods, have enabled widespread application in cancer therapy, vaccine development, and gene therapy on a large scale. This review provides a comprehensive overview of the virus biology, and several aspects of recombinant Ad vectors, as well as the development of Ad vector, are discussed. Moreover, we focus on those Ads that were used in preclinical and clinical applications including regenerative medicine, vaccine development, genome engineering, treatment of genetic diseases, and virotherapy in tumor treatment.

1. A Historical Introduction into the Discovery of the Adenovirus

Following the first isolation of adenovirus (Ad) from human adenoid tissues in 1953 [1], it has been established that Ad can infect a wide range of vertebrates [2,3]. Due to the observation of its cytopathogenic effects in adenoid and tonsil tissues during prolonged cultivation, Ad was initially termed as an ‘adenoid degeneration agent’ abbreviated as an A.D. agent [1]. A year later, during an epidemic at Fort Leonard Wood, a microbial agent was identified in throat washings from patients with primary atypic pneumonia. Similar to the earlier cases, it was observed that this unidentified agent caused cytopathogenic changes in human cell cultures and was designated as the ‘respiratory illness (RI) agent’ [4,5].
In 1954, a small infectious outbreak of febrile pharyngitis with conjunctivitis was reported from a hospital, leading to the proposal of the term ‘adenoidal-pharyngeal-conjunctival agent’ [6]. Furthermore, in July 1956, an article published in the journal Science, highlighted the necessity of assigning a common name to this group of predominantly respiratory viruses. However, institutions worldwide participated in this discussion and, during a meeting on 25 May 1956 in New York, USA, it was decided to adopt the term ‘Adenovirus group’ [7].

2. Adenovirus Biology and Clinical Presentation

Adenovirus belongs to the family of Adenoviridae [8,9,10], specifically within the genus Mastadenovirus [11,12,13]. Human Ads are categorized into seven species (A to G), comprising 114 identified types (http://hadvwg.gmu.edu/ (accessed on 22 January 2024, [14,15,16]). These non-enveloped double-stranded DNA viruses present a pseudo T = 25 icosahedral capsid with a diameter of 95 nm (vertex to vertex) [17]. The virion consists of three major proteins, i.e., hexon, penton base, and fiber [18,19].
The hexon, which represents approximately 60% of the virion mass, is a highly conserved structure throughout the Adenoviridae family, composed of more than 900 amino acids [20,21,22]. The icosahedral capsid features a penton complex at its 12 vertices, comprised the penton base and fiber [18,23,24,25,26]. The penton base is a pentameric structure that fills the space at each vertex left by the five peri-pentonal hexons [22,27,28], while the fiber is an elongated protein divided into an N-terminal ‘shaft’ anchored on the capsid and a C-terminal knob that interacts with receptor structures [21,29,30]. Minor capsid components such as proteins IIIa, VI, VIII, and IX contribute to the capsid structure and stability [31,32]. Within the capsid, there are basic proteins (protein V, VII, and µ) that are packaged within the virion together with the genome [33,34,35,36].
Adenovirus utilizes several receptors for cell entry, including Coxsackie and adenovirus receptor (CAR) [37,38,39,40], CD46 [11,41], desmoglein-2 (Dsg-2) [13] Glycans GD1a [42] and polysialic acid [43], etc. Notably, Ad binding to soluble factor IX [44] and X [45,46,47] may facilitate cell entry.
The linear double-stranded DNA genome of Ads has a length of ~34 to 36 kilo-base-pairs (kb). At the ends, the Ad genome is flanked by highly conserved sequences defined as ‘inverted terminal repeats’ (ITRs). Downstream of the left ITR, a DNA sequence serving as a packaging signal (Ψ) is located, interacting with other viral proteins to encapsidate the Ad genome [48,49,50]. The coding region of the Ad genome contains four transcription units (E1, E2, E3, and E4) that encode early genes, as well as late transcription units (L1, L2, L3, L4, and L5) that are responsible for encoding late proteins [33,51,52,53,54,55,56,57,58,59]. The E1, E2, and E4 transcription units regulate the transcription and translation of late genes, which are crucial for the adenoviral replication [60]. In the region E1, two transcription units can be identified [61]. E1A proteins play an important role during initial steps of viral infection as they stimulate the transcription of viral genes [62,63]. Transcription unit E1A inhibits the transcription of certain cellular genes by inhibiting the homo- or hetero-oligomerization of p53 [64]. E1B assists with E1A to redirect the function of the host cell and suppresses apoptosis [65,66], thereby facilitating viral replication, and also contributes to the transformation of the host cell [67]. The E2 region can be divided into two segments: the proximal segment E2A and the distal segment E2B. Both segments encode proteins which are relevant for the replication of the viral genome [68]. The E3 region plays a critical role in the pathogenesis of the disease process by inhibiting both the specific immunity (i.e., cytotoxic T lymphocytes and CTL) and the innate immune response (i.e., tumor necrosis factor (TNF)-α) [69,70]. The products of the E3 region are not relevant for virus replication but can mediate the escape from the immune system of the host through inhibition of cytokine production, alteration of antigen presentation, and cellular apoptosis [71,72]. In addition, the E4 region encodes multiple proteins, that play an important role in maintaining the stability of viral RNAs during the later stages of infection [73]. The L1 to L5 transcription units code for the major structural proteins of Ad, including the major capsid proteins hexon, fiber, and penton [74,75,76].
Furthermore, Ad is an obligate intracellular pathogen and needs the replicative and transcriptional apparatus of the host cell to enable replication and to continue its life cycle to form new viral particles [77,78]. As shown in Figure 1, following the binding of the adenoviral knob with the host cell surface receptors, the interaction between a highly conserved RGD motif penton base protein and an activated status of cellular integrins, such as αvβ3/αvβ5, may promote the Ad capsid internalization through clathrin-mediated endocytosis [79,80], while the proteins ‘fiber’ remain on the surface of the host cell [19]. There are also other internalization pathways, including lipid rafts, caveolin-mediated endocytosis, and macro-pinocytosis, that play a secondary role and are dependent on the type of the host cell [81]. During internalization, the Ad capsid vertex region (penton base, fiber, and peri-pentonal exons) disassembles and releases various proteins including the lytic protein VI, which mediate the vesicular membrane disruption [82] and the subsequent access of the capsid to the cytosol. Once in the cytosol, the leaky Ad capsid interacts with the dynein of the microtubular system of the cell. Subsequently, this interaction facilitates the passive transport of the Ad to the nuclear pore complex (NPC) and the hexon protein of Ad binds to the nucleoprotein Nup214 of NPC. The temporal interval between the initial interaction of Ad with surface receptors and the binding with NPC proteins is brief, lasting less than an h [82,83,84,85,86].
The interaction of Ad with NPC initiates a series of events, including priming, unlocking, and disruption of the virion. Ubiquitination of the Ad protein V results in the detachment of the adenoviral DNA from the protein complex, allowing the viral DNA to be transported into the nucleus [87,88]. This interaction between viral DNA and cellular replication promoters leads to the formation of viral replication compartments (VRCs). Importantly, the life cycle of Ad occurs extrachromosomally [89].
The assembly of new viral particles takes place within the nucleus and is synchronized with viral DNA synthesis [90] to facilitate adenoviral packaging [91]. The last step in the Ad life cycle involves the lysis of the host cell, which leads to the release of new virus particles. The mechanisms underlying lysis can vary and depend on both the specific virus and the characteristics of the host cell. For instance, Ad5 induces lysis through the overexpression of the adenovirus death protein (ADP) [92,93].
Figure 1. The life cycle of adenovirus. (1) Adenovirus (Ad) infection begins with the binding of Ad fiber to cellular receptors (e.g., CAR and CD46 or Desmoglein-2). (2) This binding activates the internalization of Ad through endocytosis. (3) During cytoplasmic transport, Ad presents proteins, which, in association with vesicular acidification, allow the release of the virion (4) near the nucleus. (5) The interaction with the nuclear pore complex facilitates the transport of the viral genome into the nucleus. In addition, the replication of Ad is divided into two stages: the early phase (6), during which genes are expressed that code for transcription factors. These factors regulate the expression of other viral genes as well as host cell genes, enabling the completion of the Ad life cycle. Subsequently, viral DNA replication (8) starts along with the expression of proteins encoded by late genes (9). This coordinated activity allows the assembly of new viral particles (10) and new viral particles are released through cell lysis (11). A detailed signaling cascade through virus entry is not presented in detail to preserve the overview. This figure was adapted from [89].
Figure 1. The life cycle of adenovirus. (1) Adenovirus (Ad) infection begins with the binding of Ad fiber to cellular receptors (e.g., CAR and CD46 or Desmoglein-2). (2) This binding activates the internalization of Ad through endocytosis. (3) During cytoplasmic transport, Ad presents proteins, which, in association with vesicular acidification, allow the release of the virion (4) near the nucleus. (5) The interaction with the nuclear pore complex facilitates the transport of the viral genome into the nucleus. In addition, the replication of Ad is divided into two stages: the early phase (6), during which genes are expressed that code for transcription factors. These factors regulate the expression of other viral genes as well as host cell genes, enabling the completion of the Ad life cycle. Subsequently, viral DNA replication (8) starts along with the expression of proteins encoded by late genes (9). This coordinated activity allows the assembly of new viral particles (10) and new viral particles are released through cell lysis (11). A detailed signaling cascade through virus entry is not presented in detail to preserve the overview. This figure was adapted from [89].
Viruses 16 00377 g001
Regarding clinical manifestations, Ad infection is highly prevalent and exhibits a global distribution. Specifically, Ad can cause acute infections in the host [94] and can also lead to viral persistence [95]. It can be endemic, particularly in the pediatric population, where many Ad infections are diagnosed in children [96,97]. Certain populations, such as military recruits [98,99], patients with hematopoietic stem cell transplantation (especially allogeneic transplantation) [100], patients undergoing solid organ transplantation [101], individuals with congenital immunodeficiency [102], and those with acquired immunodeficiency [103], are more susceptible to Ad infections [97], associated with increased morbidity and mortality rates [101].
Moreover, Ad species D and E are predominantly linked to the development of keratoconjunctivitis [97], while species A and B can lead to respiratory, gastrointestinal, and urinary diseases [104,105]. Species C Ads are particularly associated with respiratory symptoms and they are the serotype most frequently isolated in children with severe acute respiratory infections [106]. Ad infections elicit both innate and adaptive immune responses. The initial response against Ad infection involves the release of different chemokines and cytokines, including interferon gamma (IFN-γ), TNF, interleukin (IL)-1, IL-2, and macrophage inflammatory proteins from non-antigen-specific cells, including dendritic cells and macrophages. In addition, natural killer (NK) cells are recruited and activated to restrict viral amplification [107,108]. Simultaneously, an antigen-specific response occurs, involving antigen-presenting cells (APCs) that present adenoviral antigens through major histocompatibility, which results in the release of cytokines by APCs, inducing T helper (TH1) (cellular) responses that activate cytotoxic processes of CD8+ T cells (through IL-2 and IFN-γ secretion), or TH2 (humoral) responses that stimulate the production of antigen-specific antibodies against adenoviral antigens (through IL-4 release) [109]. The cellular and humoral response of the host act to constrain virus infection and replication.
Moreover, Ad can also lead to latent infection and persist in various tissues, such as myocardiocytes [110], or in lymphatic niches (e.g., tonsils, adenoids, and intestinal lymphatic tissue) [111,112]. This persistence of Ad leads to the question of whether latent Ad infection may contribute to cancer development. Although Ad genome sequences are detected in some malignant tumors (e.g., cell lymphoma and glioblastoma [113,114]), studies have not yet demonstrated a direct relationship between Ad infection and tumorigenesis (summarized in [115]).

3. Adenovirus as a Vector and Its Production on a Large Scale

The intrinsic properties of Ads, including their episomal nature, minimize the risk for undesired insertional mutagenesis, eliminating the potential of germ-line transmission [116]. The Ad genome is well known and easily modifiable [117], allowing the production of recombinant replication-competent [118] and replication-defective vectors [119]. With a wide tropism and high transduction capacity for both replicating and non-replicating cell types [120], Ads are excellent gene delivery tools due to their simple manipulation of their genome [121,122].
Another advantage of Ads is their strong immunogenicity [123], making them suitable for vaccine development. Figure 2 illustrates the clinical application of adenoviral vectors. So far, there are three generations of adenoviral vectors. The first-generation vector type presents the replication-defective (or incompetent) Ads, which serve as vectors in vaccine production or gene therapy. Replication-defective Ads imply that the virus cannot proliferate uncontrollably or assemble into infectious particles [119]. In these first-generation Ad vectors, the E1 region typically contains a partial or complete deletion, resulting in the loss of expression of the essential replication proteins E1A and E1B. The deletion of E1 sequences in first-generation Ad vector transgenes allows the insertion of transgenes into the E1 region. To further increase the transgene insertion capacity, the E3 region can also be deleted, allowing the insertion of transgenes of up to 8.2 kb. However, the limited duration of cargo and leaky expression of viral genes leading to immune responses are the challenges associated with first-generation Ad vectors [124].
To enhance the capacity of the vector for transporting foreign DNA (up to 14 kb), additional deletions in the E2 and E4 regions, beyond those in the E1 and E3 deletions, were introduced, characterizing the second-generation of vectors [125,126]. These vectors exhibit reduced immunogenicity compared to first-generation vectors and a lower likelihood of generating replication-competent Ads (RCA). However, the production of Ad vectors can also be associated with issues, such as the difficulty in establishing stable cell lines for production. This complexity is inherent to the second-generation vector and resulted in low viral titers and reduced gene expression [127].
To address these challenges associated with first- and second-generation vectors, the majority of adenoviral coding regions for viral proteins, excluding cis-acting elements required for vector genome replication, such as ITR sequences and encapsidation, were deleted and replaced by foreign DNA and transgenes of interest [127,128,129].
These modifications have significantly enhanced the transgene carrying capacity to up to 36 kb and have reduced the immunogenicity of these vectors. Consequently, they are safer for patient applications, and the duration of transgene expression can be extended. This development represents the third generation of Ad vectors, the so-called gutless, gene-deleted, or high-capacity adenoviral vectors [122]. The deletion of viral protein-coding sequences in these viruses makes them dependent on a helper virus (HV). Therefore, they are also characterized as helper-dependent Ads (HD-Ad). The HV must be present within the vector-producing cell, providing the necessary proteins for replication and packaging of the HD-Ad genome to the HD-Ad [128,129,130].
Figure 2. Versatile clinical applications of adenoviral vectors in vaccine design, oncolytic and gene therapy. Adenovirus-mediated gene delivery can be used for vaccine development (as Ad can deliver genes coding for immunogenic proteins into the host, which trigger the development of immunity in the host) and for oncolytic therapy (targeted therapy, which consists of administering recombinant Ads, leading to tumor cell-specific replication of recombinant Ad only in tumor cells and their consequent lysis) [131,132,133].
Figure 2. Versatile clinical applications of adenoviral vectors in vaccine design, oncolytic and gene therapy. Adenovirus-mediated gene delivery can be used for vaccine development (as Ad can deliver genes coding for immunogenic proteins into the host, which trigger the development of immunity in the host) and for oncolytic therapy (targeted therapy, which consists of administering recombinant Ads, leading to tumor cell-specific replication of recombinant Ad only in tumor cells and their consequent lysis) [131,132,133].
Viruses 16 00377 g002
For the construction of Ad vectors, there are currently two different methods in use. The direct insertion of the foreign gene into the viral genome entails using a restriction enzyme [123] and homologous recombination between two plasmids (genomic and shuttle plasmids) either in a permissive cell line or in bacteria [134,135]. The first method, also defined as the in vitro ligation method, involves the direct insertion of the gene of interest into the Ad genome and is commonly used when inserting large DNA fragments [136,137]. On the other hand, homologous recombination is the most widely used and efficient method for the production of viral vectors. It involves the co-transformation of Escherichia coli such as BJ5183 that expresses nucleases [138] and thus is capable of homologous recombination. The thus obtained recombinant Ad (with the gene of interest in its genome) can be used for the transfection of producer cells [139]. As described, adenoviral vectors feature deletions in the E1 regions (first-generation), E2 and/or E4 (second-generation), or all coding regions (third-generation) to allow viral replication and production. To compensate for these deletions, modified cell lines are engineered to express the products of the missing regions of the adenoviral genome. For the generation of first-generation Ad vectors, human embryonic kidney (HEK) 293 cells were created by inserting the E1a and E1b sequences into their chromosome 19 [140,141]. Furthermore, the 911 cell line was generated based on the insertion of an E1 sequence of Ad5 into human embryonic retinoblasts (HER) cells [142] and lung adenocarcinoma-derived A549 cells were also used as producer cells after the insertion of E1 sequences [143]. The potential generation of RCA and a poor production of viruses are the common problems with the use of these cell lines [144]. For the second-generation vectors, cell lines deriving from HEK293, 911, and A549 cells with inserted E2 and/or E4 are used [145,146]. However, a problem arises due to the toxicity of the E2a and E4 proteins to the host cell [147]. The 293-C2 cells, which express E2a, have a viral production rate 30 times lower than the wild-type E2 vector [148]. Consequently, an E2T cell line was generated with intracellular expression levels of E2A regulated by tetracycline-controlling elements, reducing toxicity and enabling a yield similar to the wild-type E2 vector [146]. The VL2-20 and VK10-9 cells, containing the entire E4 sequence regulated by the dexamethasone-inducible MMTV promoter, are used for the production of E4-deleted adenoviral vectors [145]. For the production of third-generation vectors, not only is a suitable producer cell line required but a complementing Ad vector (helper vector) also provides packaging proteins to the viral particles being produced [149]. Through the expression of a Cre recombinase in the producer cells, encapsidation of the helper virus DNA can be prevented, ensuring that only the viral vectors can be produced [150,151].
Through the use of producer cells, the primary stock of the viral vector is created and subsequently amplified to obtain an adequate quantity of Ad vector in clinical and preclinical applications [152]. To be used in clinical settings, the Ad vector must be purified using methods such as cesium chloride density gradients [153], combined with ultracentrifugation [154], which is the most commonly used method to separate Ad vectors from other cellular debris. Using bioreactors containing continuous cell lines such as PER.C6 [151,155], a large scale production of Ad vector can be carried out. The subsequent purification involves chromatography-tandem ultracentrifugation or sequential chromatography [156].

4. Clinical Application

One of the first clinical applications of the Ad dates back to the 1950s, when live Ad4 and Ad7 were used as oral vaccines in North American military recruits to prevent respiratory disease (summarized in [157]). In 1989, the first in vivo gene transfer was demonstrated by a recombinant replication-deficient Ad with deleted E1 and E3 regions coding for β-galactosidase [158]. It was shown that Ad was effective in transferring genes in patients with hereditary α1-antitrypsin deficiency and cystic fibrosis [159]. Today, the potential clinical applications of Ad-mediated gene delivery range from regenerative medicine, vaccine development, and anticancer therapy up to gene therapy for monogenic diseases [124]. The following chapters provide an overview of these applications.

4.1. Regenerative Medicine

In 2006, Dr. Heather Greenwood defined regenerative medicine as an emerging interdisciplinary field of research and clinical applications. This field concentrates on the repair, replacement, or regeneration of cells, tissues, or organs, aiming to restore impaired function due to several causes, including congenital defects, diseases, trauma, and aging [160,161].
Within regenerative medicine, viral vectors find applications through two primary delivery approaches: in vivo [158] or ex vivo delivery [162]. In the first case, the viral vector is directly introduced into the patient to reach the targeted site of action. In the latter, the viral vector is inserted into target cells in a laboratory setting and treated cells are subsequently inserted to the site of damage [163].
In the context of Ad-based gene therapy for bone regeneration, the gene encoding the osteoinductive bone morphogenetic protein 2 (BMP2) plays a role in osteogenesis. Ex vivo approaches use the recombinant Ad-BMP2 to transduce mesenchymal stem cells, which are then implanted into bone defects within a matrix consisting of polylactide granules (which serve as a depot for genetic constructs or matrices for cell attachment) and with platelet-rich plasma as a source of growth factors and a binding gel. This ex vivo approach demonstrated a higher osteo-inductive effect compared to in vivo approaches [163].
In a pig infarction model, myocardial injection of recombinant Ad-VEGF121 improved myocardial perfusion and wall thickening after four weeks [164]. Similarly, intra-myocardial administration in the peri-infarct area of recombinant Ad-VEGF165 induced angiogenesis in the border zone of myocardial infarction [165]. In addition, endovascular intra-myocardial administration of Ad-VEGF-DΔNΔC increased perfusion and the ejection fraction in the infarct border zone in a porcine myocardial infarction model [166].

4.2. Adenoviral Vector Vaccine

The use of an adenoviral vector in vaccine design offers several advantages. Adenovirus can activate the innate immune system without the need for adjuvant substances [167]. The interaction of Ad with both surface and intracellular host proteins, such as integrins, pattern-recognition receptors, toll-like receptors (TLR), such as TLR-2, TLR-4, and TLR-9 [168], lactoferrin [169], and MyD88 [170], moderately stimulates an innate immune reaction [167]. This moderate activation of the innate immune response is appropriate to activate adaptive immune responses to transgene Ad products without causing severe side effects due to the excessive release of pro-inflammatory cytokines [171,172].
The adaptive immune response induced by Ad includes not only the production of antibodies but also a robust activation of CTL. Ad vector can transduce both immune and non-immune cells [173]. When transgene expression occurs in non-immune cells, immunogenic products are released from the cells [172]. The uptake of these products by APCs leads to the production of various isotypes and subclasses of specific antibodies [174,175]. Moreover, intramuscular administration can also stimulate moderate levels of mucosal immunity activity [172,176] through the induction of T helper 17 cells, which migrate to the gut mucosa and stimulate the proliferation of mucosal antigen-specific CTL [177].
High seroprevalence to Ad5 in the general population represents the strongest obstacle for conventional Ad-based vector vaccines [178,179]. Neutralizing antibodies against Ad5 are specific for hexon, fiber, and rarely for penton base proteins. Strategies to overcome these challenges include the use of serotypes with low seroprevalence, such as Ad11, 26, 35, 48; the application of recombinant Ad with modified hexon proteins [180,181]; and the utilization of non-human Ad vectors [182]. Moreover, strategies to hide the surface antigens of Ad have been developed, including the use of polyethylene glycol [183] coats, liposome-encapsulated vectors [184], and vesicle encapsulation techniques [185].
One of the most notable Ads in vaccine development is during the previous COVID-19 pandemic [186]. Vaccines such as ChAdOX1 nCoV-19 [187,188,189], Sputnik V [190], and Ad5-nCOV [191] are Ad-based vector vaccines expressing the full-length spike protein of SARS-CoV-2, while Ad26.COV2-S [192] codes for a recombinant spike protein. Other Ad-based vaccines include Ad26.ZEBOV (encoding glycoproteins from Ebola, Sudan, Marburg, and Tai Forest viruses nucleoprotein) [193], Ad26.ZIKV.001 (inducing neutralizing antibodies against Zika virus) [194], Ad26.Mos.HIV (an Ad26-based vaccine) [195,196], and ChAdOx1 (a replication-incompetent Ad5-based vaccine expressing influenza virus antigens nucleoprotein (NP) and matrix protein-1 (M1) [172,197]. Recent clinical trials about the application of adenoviral vectors in vaccine production are listed in Table 1.

4.3. Adenoviral Vectors and Gene Editing

Adenovirus has also demonstrated a versatile role in the field of genome editing techniques. The clustered regularly interspaced short palindromic repeat (CRISPR) and its accompanying protein (Cas9) is a highly efficient and precise technique enabling scientists to modify the DNA of an organism [198]. This methodology includes two important components: a guide RNA to match the target gene and an endonuclease, which induces a double-stranded DNA break [199]. Although adeno-associated viral vectors (AAVs) play an important role [200], adenoviral vectors can also be used to transport all the components of the system CRISPR/Cas9 [201], facilitating an in vivo-knock-in approach in models [202].
Furthermore, efforts to perform gene editing in hepatocytes have been undertaken using the CRISPR/Cpf1 system [203]. However, the widespread seropositivity among the general population for Ad represents a significant limitation in the use of viral vectors in this field [124].

4.4. Oncolytic Adenovirus and Cancer Therapy

The use of recombinant Ad as an oncolytic vector represents a novel strategy in cancer therapy, increasingly recognized in clinical trials since 2001, when oncolytic Ad (OAd) Onyx-015 was employed as a therapeutic intervention for advanced pancreatic cancer [204,205]. A critical requirement for ensuring the safety and efficacy of this therapeutic approach is the replication of OAd exclusively in tumor cells while healthy tissue does not perfectly enable vector replication [206]. The prototype of Ad-mediated virotherapy was ONYX-015 [207], an OAd characterized by a deletion in the region coding for E1b protein. Typically, the E1b protein binds to and inactivates p53 to promote viral replication. However, OAd lacking the E1b gene can replicate in cancer cells that are deficient in p53 expression [207]. Another example is demonstrated by an OAd carrying inactivating deletions in the region coding for the E1a protein [208]. Normally, E1a interacts with the retinoblastoma (Rb) protein of the host cell to neutralize its inhibitory effect on E2F, a transcriptional activator that initiates the S-phase of the cell cycle [209]. Consequently, OAd vectors with deleted E1a can replicate in cancer cells, where the expression of the Rb protein is suppressed [210]. However, the absence of OAd receptors on the surface of certain tumor cells limits its applicability [204].
The use of oncolytic Ad5 from species C is limited to cancer cells with the CAR on their cell membrane. To address this limitation, utilizing group B Ad as oncolytic vectors, which present a higher affinity for the CD46 receptor, offers advantages. For instance, oncolytic Ad35 demonstrates a 100-fold higher activity than oncolytic Ad5 against breast cancer [211]. Similarly, Ad5 has a low efficacy in hematological malignancies, due to low expression of CAR on blood cells [212]. Despite the good affinity of oncolytic Ads of species D for CD46 and salicylic acid [213], their short fiber reduces the binding affinity to the Ad receptor, making them more suitable for vaccine development [214]. In addition, oncolytic Ads from groups E, F, and G, which primarily interact with CAR receptor, have not been successful in this therapeutic context [206].
The insertion of a transgene coding for adjuvant factors into the Ads genome can improve their therapeutic efficacy, acting synergistically with their oncolytic action [215]. Transgenes can be also inserted into the E3 region [216]. Either human cytomegalovirus promoters [217] or heat shock protein promoters [218] are designed to express the transgene. A further approach is to insert the transgene into a late gene, which has shown an advantage in terms of expression level and specificity [219]. Notably, armed Ad vectors carrying molecules such as GM-CSF [220], IL-12 [221], IL-15 [222], INF-γ [223], and TNF-α [224] have demonstrated significant success in clinical trials. A recent advancement involves the combination of armed OAd with chimeric antigen T-cell (CAR-T) therapy [225], as well as chimeric antigen receptor-NK cell therapy [206,226,227]. These applications show the potential for armed OAd to play a pivotal role in the future landscape of cancer treatment. Table 2 lists a few clinical trials of oncolytic adenoviral vectors.

5. Conclusions

Gene therapy using adenoviral vectors has been recognized as a promising therapeutic option for several clinical applications. The ongoing monitoring and analysis of outcomes among patients participating in clinical trials remain important for a thorough understanding of the long-term efficacy and safety of recombinant Ad vectors. The episomal nature of the adenoviral genome represents an important advantage over the commonly used recombinant lentiviruses in clinical settings due to the high transduction efficiency of Ads. However, despite these advantages, challenges remain. The widespread seroprevalence of Ads in the population presents issues that require careful consideration. In cancer therapy, where synergistic advantages have been observed, this includes the combination of armed oncolytic viruses and CAR-NK cell therapy [226]. Addressing these challenges and integrating therapeutic strategies at multiple levels is promising for advancing the field of gene therapy.

Author Contributions

Conceptualization, L.S., E.E.-S., S.C.T., M.P., A.E. and M.A.; methodology, L.S. and M.A.; software, L.S. and M.A.; validation, L.S., E.E.-S., A.E. and M.A.; investigation, L.S. and M.A.; resources, L.S.; writing—original draft preparation, L.S., E.E.-S. and M.A.; writing—review and editing, all authors; visualization, L.S.; supervision, A.E. and M.A.; project administration, A.E. and M.A.; funding acquisition, L.S., A.E. and M.A. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by internal research funding (UW/H-IFF-2023-62 and IFF2023-77) from the Witten/Herdecke University, Witten, Germany, and DFG grant EH 192/5-3.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Figure 1 and Figure 2 were created using https://www.biorender.com/ (accessed on 15 September 2023). The clinical trials, listed in Table 1 and Table 2, were found through the website https://clinicaltrials.gov/ (accessed on 15 September 2023).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Rowe, W.P.; Huebner, R.J.; Gilmore, L.K.; Parrott, R.H.; Ward, T.G. Isolation of a cytopathogenic agent from human adenoids undergoing spontaneous degeneration in tissue culture. Proc. Soc. Exp. Biol. Med. 1953, 84, 570–573. [Google Scholar] [CrossRef]
  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. Pereira, H.G. Adenoviruses of man and animals. Dev. Biol. Stand. 1975, 28, 28–41. [Google Scholar]
  4. Gray, G.C.; Callahan, J.D.; Hawksworth, A.W.; Fisher, C.A.; Gaydos, J.C. Respiratory diseases among U.S. military personnel: Countering emerging threats. Emerg. Infect. Dis. 1999, 5, 379–385. [Google Scholar] [CrossRef]
  5. Hilleman, M.R.; Werner, J.H. Recovery of new agent from patients with acute respiratory illness. Proc. Soc. Exp. Biol. Med. 1954, 85, 183–188. [Google Scholar] [CrossRef]
  6. Huebner, R.J.; Rowe, W.P.; Ward, T.G.; Parrott, R.H.; Bell, J.A. Adenoidal-pharyngeal-conjunctival agents: A newly recognized group of common viruses of the respiratory system. N. Engl. J. Med. 1954, 251, 1077–1086. [Google Scholar] [CrossRef]
  7. Enders, J.F.; Bell, J.A.; Dingle, J.H.; Francis, T.; Hilleman, M.R.; Huebner, R.J.; Payne, A.M. Adenoviruses: Group name proposed for new respiratory-tract viruses. Science 1956, 124, 119–120. [Google Scholar] [CrossRef]
  8. Shenk, T. Adenoviridae: The Viruses and Their Replication, Fields Virology, 3rd ed.; Lippincott-Raven Publishers: Philadelphia, PA, USA, 1996; Volume 2. [Google Scholar]
  9. Roelvink, P.W.; Mi Lee, G.; Einfeld, D.A.; Kovesdi, I.; Wickham, T.J. Identification of a conserved receptor-binding site on the fiber proteins of CAR-recognizing adenoviridae. Science 1999, 286, 1568–1571. [Google Scholar] [CrossRef]
  10. Smith, J.G.; Wiethoff, C.M.; Stewart, P.L.; Nemerow, G.R. Adenovirus. Curr. Top. Microbiol. Immunol. 2010, 343, 195–224. [Google Scholar] [CrossRef]
  11. Gaggar, A.; Shayakhmetov, D.M.; Lieber, A. CD46 is a cellular receptor for group B adenoviruses. Nat. Med. 2003, 9, 1408–1412. [Google Scholar] [CrossRef]
  12. Hanaoka, N.; Hazama, M.; Fukushima, K.; Fujimoto, T. Sensitivity of Human Mastadenovirus, the Causal Agent of Pharyngoconjunctival Fever, Epidemic Keratoconjunctivitis, and Hemorrhagic Cystitis in Immunocompromised Individuals, to Brincidofovir. Microbiol. Spectr. 2022, 10, e0156921. [Google Scholar] [CrossRef]
  13. Wang, H.; Li, Z.-Y.; Liu, Y.; Persson, J.; Beyer, I.; Möller, T.; Koyuncu, D.; Drescher, M.R.; Strauss, R.; Zhang, X.-B.; et al. Desmoglein 2 is a receptor for adenovirus serotypes 3, 7, 11 and 14. Nat. Med. 2011, 17, 96–104. [Google Scholar] [CrossRef]
  14. Benkő, M.; Aoki, K.; Arnberg, N.; Davison, A.J.; Echavarría, M.; Hess, M.; Jones, M.S.; Kaján, G.L.; Kajon, A.E.; Mittal, S.K.; et al. ICTV Virus Taxonomy Profile: Adenoviridae 2022. J. Gen. Virol. 2022, 103, 001721. [Google Scholar] [CrossRef]
  15. Crenshaw, B.J.; Jones, L.B.; Bell, C.R.; Kumar, S.; Matthews, Q.L. Perspective on Adenoviruses: Epidemiology, Pathogenicity, and Gene Therapy. Biomedicines 2019, 7, 61. [Google Scholar] [CrossRef] [PubMed]
  16. Wang, Y.; Li, Y.; Lu, R.; Zhao, Y.; Xie, Z.; Shen, J.; Tan, W. Phylogenetic evidence for intratypic recombinant events in a novel human adenovirus C that causes severe acute respiratory infection in children. Sci. Rep. 2016, 6, 23014. [Google Scholar] [CrossRef]
  17. Robinson, C.M.; Singh, G.; Lee, J.Y.; Dehghan, S.; Rajaiya, J.; Liu, E.B.; Yousuf, M.A.; Betensky, R.A.; Jones, M.S.; Dyer, D.W.; et al. Molecular evolution of human adenoviruses. Sci. Rep. 2013, 3, 1812. [Google Scholar] [CrossRef]
  18. van Oostrum, J.; Burnett, R.M. Molecular composition of the adenovirus type 2 virion. J. Virol. 1985, 56, 439–448. [Google Scholar] [CrossRef]
  19. Nemerow, G.R.; Stewart, P.L.; Reddy, V.S. Structure of human adenovirus. Curr. Opin. Virol. 2012, 2, 115–121. [Google Scholar] [CrossRef] [PubMed]
  20. Rux, J.J.; Kuser, P.R.; Burnett, R.M. Structural and phylogenetic analysis of adenovirus hexons by use of high-resolution x-ray crystallographic, molecular modeling, and sequence-based methods. J. Virol. 2003, 77, 9553–9566. [Google Scholar] [CrossRef]
  21. Burnett, R.M.; Grütter, M.G.; White, J.L. The structure of the adenovirus capsid. I. An envelope model of hexon at 6 A resolution. J. Mol. Biol. 1985, 185, 105–123. [Google Scholar] [CrossRef] [PubMed]
  22. Condezo, G.N.; Martín-González, N.; Pérez-Illana, M.; Hernando-Pérez, M.; Gallardo, J.; San Martín, C. Adenoviruses (Adenoviridae) and their structural relatives. In Encyclopedia of Virology; Elsevier: Amsterdam, The Netherlands, 2021; pp. 329–344. [Google Scholar]
  23. Crawford-Miksza, L.; Schnurr, D.P. Analysis of 15 adenovirus hexon proteins reveals the location and structure of seven hypervariable regions containing serotype-specific residues. J. Virol. 1996, 70, 1836–1844. [Google Scholar] [CrossRef]
  24. Sevvana, M.; Klose, T.; Rossmann, M.G. Principles of Virus Structure. In Encyclopedia of Virology; Elsevier Academic Press: Amsterdam, The Netherlands, 2021; pp. 257–277. ISBN 9780128145166. [Google Scholar]
  25. Caspar, D.L.; Klug, A. Physical principles in the construction of regular viruses. Cold Spring Harb. Symp. Quant. Biol. 1962, 27, 1–24. [Google Scholar] [CrossRef]
  26. Liu, H.; Jin, L.; Koh, S.B.S.; Atanasov, I.; Schein, S.; Wu, L.; Zhou, Z.H. Atomic structure of human adenovirus by cryo-EM reveals interactions among protein networks. Science 2010, 329, 1038–1043. [Google Scholar] [CrossRef]
  27. Zubieta, C.; Schoehn, G.; Chroboczek, J.; Cusack, S. The structure of the human adenovirus 2 penton. Mol. Cell 2005, 17, 121–135. [Google Scholar] [CrossRef] [PubMed]
  28. Stewart, P.L.; Fuller, S.D.; Burnett, R.M. Difference imaging of adenovirus: Bridging the resolution gap between X-ray crystallography and electron microscopy. EMBO J. 1993, 12, 2589–2599. [Google Scholar] [CrossRef] [PubMed]
  29. Singh, A.K.; Menéndez-Conejero, R.; San Martín, C.; van Raaij, M.J. Crystal structure of the fibre head domain of the Atadenovirus Snake Adenovirus 1. PLoS ONE 2014, 9, e114373. [Google Scholar] [CrossRef] [PubMed]
  30. Chroboczek, J.; Ruigrok, R.W.; Cusack, S. Adenovirus fiber. Curr. Top. Microbiol. Immunol. 1995, 199 Pt 1, 163–200. [Google Scholar] [CrossRef]
  31. Gallardo, J.; Pérez-Illana, M.; Martín-González, N.; San Martín, C. Adenovirus Structure: What Is New? Int. J. Mol. Sci. 2021, 22, 5240. [Google Scholar] [CrossRef]
  32. Ma, H.-C.; Hearing, P. Adenovirus structural protein IIIa is involved in the serotype specificity of viral DNA packaging. J. Virol. 2011, 85, 7849–7855. [Google Scholar] [CrossRef]
  33. Davison, A.J.; Benkő, M.; Harrach, B. Genetic content and evolution of adenoviruses. J. Gen. Virol. 2003, 84, 2895–2908. [Google Scholar] [CrossRef] [PubMed]
  34. Russell, W.C. Adenoviruses: Update on structure and function. J. Gen. Virol. 2009, 90, 1–20. [Google Scholar] [CrossRef]
  35. Martín-González, N.; Gómez-González, A.; Hernando-Pérez, M.; Bauer, M.; Greber, U.F.; San Martín, C.; de Pablo, P.J. Adenovirus core protein V reinforces the capsid and enhances genome release from disrupted particles. Sci. Adv. 2023, 9, eade9910. [Google Scholar] [CrossRef]
  36. Pérez-Berná, A.J.; Marion, S.; Chichón, F.J.; Fernández, J.J.; Winkler, D.C.; Carrascosa, J.L.; Steven, A.C.; Šiber, A.; San Martín, C. Distribution of DNA-condensing protein complexes in the adenovirus core. Nucleic Acids Res. 2015, 43, 4274–4283. [Google Scholar] [CrossRef]
  37. Bergelson, J.M.; Cunningham, J.A.; Droguett, G.; Kurt-Jones, E.A.; Krithivas, A.; Hong, J.S.; Horwitz, M.S.; Crowell, R.L.; Finberg, R.W. Isolation of a common receptor for Coxsackie B viruses and adenoviruses 2 and 5. Science 1997, 275, 1320–1323. [Google Scholar] [CrossRef]
  38. Roelvink, P.W.; Lizonova, A.; Lee, J.G.; Li, Y.; Bergelson, J.M.; Finberg, R.W.; Brough, D.E.; Kovesdi, I.; Wickham, T.J. The coxsackievirus-adenovirus receptor protein can function as a cellular attachment protein for adenovirus serotypes from subgroups A, C, D, E, and F. J. Virol. 1998, 72, 7909–7915. [Google Scholar] [CrossRef]
  39. Tomko, R.P.; Xu, R.; Philipson, L. HCAR and MCAR: The human and mouse cellular receptors for subgroup C adenoviruses and group B coxsackieviruses. Proc. Natl. Acad. Sci. USA 1997, 94, 3352–3356. [Google Scholar] [CrossRef] [PubMed]
  40. Bewley, M.C.; Springer, K.; Zhang, Y.B.; Freimuth, P.; Flanagan, J.M. Structural analysis of the mechanism of adenovirus binding to its human cellular receptor, CAR. Science 1999, 286, 1579–1583. [Google Scholar] [CrossRef] [PubMed]
  41. Marttila, M.; Persson, D.; Gustafsson, D.; Liszewski, M.K.; Atkinson, J.P.; Wadell, G.; Arnberg, N. CD46 is a cellular receptor for all species B adenoviruses except types 3 and 7. J. Virol. 2005, 79, 14429–14436. [Google Scholar] [CrossRef]
  42. Nilsson, E.C.; Storm, R.J.; Bauer, J.; Johansson, S.M.C.; Lookene, A.; Ångström, J.; Hedenström, M.; Eriksson, T.L.; Frängsmyr, L.; Rinaldi, S.; et al. The GD1a glycan is a cellular receptor for adenoviruses causing epidemic keratoconjunctivitis. Nat. Med. 2011, 17, 105–109. [Google Scholar] [CrossRef]
  43. Lenman, A.; Liaci, A.M.; Liu, Y.; Frängsmyr, L.; Frank, M.; Blaum, B.S.; Chai, W.; Podgorski, I.I.; Harrach, B.; Benkő, M.; et al. Polysialic acid is a cellular receptor for human adenovirus 52. Proc. Natl. Acad. Sci. USA 2018, 115, E4264–E4273. [Google Scholar] [CrossRef]
  44. Lenman, A.; Müller, S.; Nygren, M.I.; Frängsmyr, L.; Stehle, T.; Arnberg, N. Coagulation factor IX mediates serotype-specific binding of species A adenoviruses to host cells. J. Virol. 2011, 85, 13420–13431. [Google Scholar] [CrossRef]
  45. Arnberg, N. Adenovirus receptors: Implications for targeting of viral vectors. Trends Pharmacol. Sci. 2012, 33, 442–448. [Google Scholar] [CrossRef]
  46. Findlay, J.S.; Cook, G.P.; Blair, G.E. Blood Coagulation Factor X Exerts Differential Effects on Adenovirus Entry into Human Lymphocytes. Viruses 2018, 10, 20. [Google Scholar] [CrossRef]
  47. 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]
  48. Doszpoly, A.; Harrach, B.; LaPatra, S.; Benkő, M. Unconventional gene arrangement and content revealed by full genome analysis of the white sturgeon adenovirus, the single member of the genus Ichtadenovirus. Infect. Genet. Evol. 2019, 75, 103976. [Google Scholar] [CrossRef] [PubMed]
  49. Davison, A.J.; Wright, K.M.; Harrach, B. DNA sequence of frog adenovirus. J. Gen. Virol. 2000, 81, 2431–2439. [Google Scholar] [CrossRef] [PubMed]
  50. Ostapchuk, P.; Hearing, P. Control of adenovirus packaging. J. Cell. Biochem. 2005, 96, 25–35. [Google Scholar] [CrossRef] [PubMed]
  51. O’Connor, R.J.; Hearing, P. The C-terminal 70 amino acids of the adenovirus E4-ORF6/7 protein are essential and sufficient for E2F complex formation. Nucleic Acids Res. 1991, 19, 6579–6586. [Google Scholar] [CrossRef] [PubMed]
  52. Lutz, P.; Rosa-Calatrava, M.; Kedinger, C. The product of the adenovirus intermediate gene IX is a transcriptional activator. J. Virol. 1997, 71, 5102–5109. [Google Scholar] [CrossRef] [PubMed]
  53. Huang, M.M.; Hearing, P. The adenovirus early region 4 open reading frame 6/7 protein regulates the DNA binding activity of the cellular transcription factor, E2F, through a direct complex. Genes Dev. 1989, 3, 1699–1710. [Google Scholar] [CrossRef] [PubMed]
  54. Hardy, S.; Engel, D.A.; Shenk, T. An adenovirus early region 4 gene product is required for induction of the infection-specific form of cellular E2F activity. Genes Dev. 1989, 3, 1062–1074. [Google Scholar] [CrossRef]
  55. Chang, L.S.; Shenk, T. The adenovirus DNA-binding protein stimulates the rate of transcription directed by adenovirus and adeno-associated virus promoters. J. Virol. 1990, 64, 2103–2109. [Google Scholar] [CrossRef]
  56. Fessler, S.P.; Young, C.S. Control of adenovirus early gene expression during the late phase of infection. J. Virol. 1998, 72, 4049–4056. [Google Scholar] [CrossRef] [PubMed]
  57. Berk, A.J. Functions of adenovirus E1A. Cancer Surv. 1986, 5, 367–387. [Google Scholar] [PubMed]
  58. Kulanayake, S.; Tikoo, S.K. Adenovirus Core Proteins: Structure and Function. Viruses 2021, 13, 388. [Google Scholar] [CrossRef]
  59. Hall, K.; Blair Zajdel, M.E.; Blair, G.E. Unity and diversity in the human adenoviruses: Exploiting alternative entry pathways for gene therapy. Biochem. J. 2010, 431, 321–336. [Google Scholar] [CrossRef] [PubMed]
  60. Giberson, A.N.; Davidson, A.R.; Parks, R.J. Chromatin structure of adenovirus DNA throughout infection. Nucleic Acids Res. 2012, 40, 2369–2376. [Google Scholar] [CrossRef]
  61. Ishida, S.; Fujinaga, Y.; Fujinaga, K.; Sakamoto, N.; Hashimoto, S. Unusual splice sites in the E1A-E1B cotranscripts synthesized in adenovirus type 40-infected A549 cells. Arch. Virol. 1994, 139, 389–402. [Google Scholar] [CrossRef]
  62. Lillie, J.W.; Loewenstein, P.M.; Green, M.R.; Green, M. Functional domains of adenovirus type 5 E1a proteins. Cell 1987, 50, 1091–1100. [Google Scholar] [CrossRef]
  63. Avvakumov, N.; Wheeler, R.; D’Halluin, J.C.; Mymryk, J.S. Comparative sequence analysis of the largest E1A proteins of human and simian adenoviruses. J. Virol. 2002, 76, 7968–7975. [Google Scholar] [CrossRef]
  64. Steegenga, W.T.; van Laar, T.; Riteco, N.; Mandarino, A.; Shvarts, A.; van der Eb, A.J.; Jochemsen, A.G. Adenovirus E1A proteins inhibit activation of transcription by p53. Mol. Cell. Biol. 1996, 16, 2101–2109. [Google Scholar] [CrossRef] [PubMed]
  65. van den Elsen, P.; Houweling, A.; van der Eb, A. Expression of region E1b of human adenoviruses in the absence of region E1a is not sufficient for complete transformation. Virology 1983, 128, 377–390. [Google Scholar] [CrossRef] [PubMed]
  66. Boulakia, C.A.; Chen, G.; Ng, F.W.; Teodoro, J.G.; Branton, P.E.; Nicholson, D.W.; Poirier, G.G.; Shore, G.C. Bcl-2 and adenovirus E1B 19 kDA protein prevent E1A-induced processing of CPP32 and cleavage of poly(ADP-ribose) polymerase. Oncogene 1996, 12, 529–535. [Google Scholar]
  67. Hidalgo, P.; Ip, W.H.; Dobner, T.; Gonzalez, R.A. The biology of the adenovirus E1B 55K protein. FEBS Lett. 2019, 593, 3504–3517. [Google Scholar] [CrossRef] [PubMed]
  68. Caravokyri, C.; Leppard, K.N. Human adenovirus type 5 variants with sequence alterations flanking the E2A gene: Effects on E2 expression and DNA replication. Virus Genes 1996, 12, 65–75. [Google Scholar] [CrossRef] [PubMed]
  69. Deryckere, F.; Ebenau-Jehle, C.; Wold, W.S.; Burgert, H.G. Tumor necrosis factor alpha increases expression of adenovirus E3 proteins. Immunobiology 1995, 193, 186–192. [Google Scholar] [CrossRef] [PubMed]
  70. Sparer, T.E.; Tripp, R.A.; Dillehay, D.L.; Hermiston, T.W.; Wold, W.S.; Gooding, L.R. The role of human adenovirus early region 3 proteins (gp19K, 10.4K, 14.5K, and 14.7K) in a murine pneumonia model. J. Virol. 1996, 70, 2431–2439. [Google Scholar] [CrossRef]
  71. Täuber, B.; Dobner, T. Molecular regulation and biological function of adenovirus early genes: The E4 ORFs. Gene 2001, 278, 1–23. [Google Scholar] [CrossRef]
  72. Lichtenstein, D.L.; Toth, K.; Doronin, K.; Tollefson, A.E.; Wold, W.S.M. Functions and mechanisms of action of the adenovirus E3 proteins. Int. Rev. Immunol. 2004, 23, 75–111. [Google Scholar] [CrossRef]
  73. Sandler, A.B.; Ketner, G. Adenovirus early region 4 is essential for normal stability of late nuclear RNAs. J. Virol. 1989, 63, 624–630. [Google Scholar] [CrossRef]
  74. Seto, D.; Chodosh, J.; Brister, J.R.; Jones, M.S. Using the whole-genome sequence to characterize and name human adenoviruses. J. Virol. 2011, 85, 5701–5702. [Google Scholar] [CrossRef] [PubMed]
  75. Liu, L. Characterization of an upstream regulatory element of adenovirus L1 poly (A) site. Virology 2005, 337, 124–135. [Google Scholar] [CrossRef] [PubMed]
  76. Seggern, D.J.V.; Chiu, C.Y.; Fleck, S.K.; Stewart, P.L.; Nemerow, G.R. A helper-independent adenovirus vector with E1, E3, and fiber deleted: Structure and infectivity of fiberless particles. J. Virol. 1999, 73, 1601–1608. [Google Scholar] [CrossRef] [PubMed]
  77. Kremer, E.J.; Nemerow, G.R. Adenovirus tales: From the cell surface to the nuclear pore complex. PLoS Pathog. 2015, 11, e1004821. [Google Scholar] [CrossRef] [PubMed]
  78. Greber, U.F.; Flatt, J.W. Adenovirus Entry: From Infection to Immunity. Annu. Rev. Virol. 2019, 6, 177–197. [Google Scholar] [CrossRef] [PubMed]
  79. Nemerow, G.R. Cell receptors involved in adenovirus entry. Virology 2000, 274, 1–4. [Google Scholar] [CrossRef] [PubMed]
  80. Wickham, T.J.; Mathias, P.; Cheresh, D.A.; Nemerow, G.R. Integrins alpha v beta 3 and alpha v beta 5 promote adenovirus internalization but not virus attachment. Cell 1993, 73, 309–319. [Google Scholar] [CrossRef] [PubMed]
  81. Nestić, D.; Božinović, K.; Pehar, I.; Wallace, R.; Parker, A.L.; Majhen, D. The Revolving Door of Adenovirus Cell Entry: Not All Pathways Are Equal. Pharmaceutics 2021, 13, 1585. [Google Scholar] [CrossRef]
  82. Wiethoff, C.M.; Wodrich, H.; Gerace, L.; Nemerow, G.R. Adenovirus protein VI mediates membrane disruption following capsid disassembly. J. Virol. 2005, 79, 1992–2000. [Google Scholar] [CrossRef]
  83. Cassany, A.; Ragues, J.; Guan, T.; Bégu, D.; Wodrich, H.; Kann, M.; Nemerow, G.R.; Gerace, L. Nuclear import of adenovirus DNA involves direct interaction of hexon with an N-terminal domain of the nucleoporin Nup214. J. Virol. 2015, 89, 1719–1730. [Google Scholar] [CrossRef]
  84. 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]
  85. Pied, N.; Wodrich, H. Imaging the adenovirus infection cycle. FEBS Lett. 2019, 593, 3419–3448. [Google Scholar] [CrossRef]
  86. Wang, I.-H.; Burckhardt, C.J.; Yakimovich, A.; Morf, M.K.; Greber, U.F. The nuclear export factor CRM1 controls juxta-nuclear microtubule-dependent virus transport. J. Cell Sci. 2017, 130, 2185–2195. [Google Scholar] [CrossRef]
  87. Trotman, L.C.; Mosberger, N.; Fornerod, M.; Stidwill, R.P.; Greber, U.F. Import of adenovirus DNA involves the nuclear pore complex receptor CAN/Nup214 and histone H1. Nat. Cell Biol. 2001, 3, 1092–1100. [Google Scholar] [CrossRef]
  88. Greber, U.F.; Suomalainen, M. Adenovirus entry: Stability, uncoating, and nuclear import. Mol. Microbiol. 2022, 118, 309–320. [Google Scholar] [CrossRef]
  89. Charman, M.; Herrmann, C.; Weitzman, M.D. Viral and cellular interactions during adenovirus DNA replication. FEBS Lett. 2019, 593, 3531–3550. [Google Scholar] [CrossRef]
  90. Condezo, G.N.; San Martín, C. Localization of adenovirus morphogenesis players, together with visualization of assembly intermediates and failed products, favor a model where assembly and packaging occur concurrently at the periphery of the replication center. PLoS Pathog. 2017, 13, e1006320. [Google Scholar] [CrossRef]
  91. Hammarskjöld, M.L.; Winberg, G. Encapsidation of adenovirus 16 DNA is directed by a small DNA sequence at the left end of the genome. Cell 1980, 20, 787–795. [Google Scholar] [CrossRef]
  92. Murali, V.K.; Ornelles, D.A.; Gooding, L.R.; Wilms, H.T.; Huang, W.; Tollefson, A.E.; Wold, W.S.M.; Garnett-Benson, C. Adenovirus death protein (ADP) is required for lytic infection of human lymphocytes. J. Virol. 2014, 88, 903–912. [Google Scholar] [CrossRef]
  93. Doronin, K.; Toth, K.; Kuppuswamy, M.; Krajcsi, P.; Tollefson, A.E.; Wold, W.S.M. Overexpression of the ADP (E3-11.6K) protein increases cell lysis and spread of adenovirus. Virology 2003, 305, 378–387. [Google Scholar] [CrossRef]
  94. Chen, Y.; Lin, T.; Wang, C.-B.; Liang, W.-L.; Lian, G.-W.; Zanin, M.; Wong, S.-S.; Tian, X.-G.; Zhong, J.-Y.; Zhang, Y.-Y.; et al. Human adenovirus (HAdV) infection in children with acute respiratory tract infections in Guangzhou, China, 2010-2021: A molecular epidemiology study. World J. Pediatr. 2022, 18, 545–552. [Google Scholar] [CrossRef]
  95. Radke, J.R.; Cook, J.L. Human adenovirus infections: Update and consideration of mechanisms of viral persistence. Curr. Opin. Infect. Dis. 2018, 31, 251–256. [Google Scholar] [CrossRef] [PubMed]
  96. Mitchell, L.S.; Taylor, B.; Reimels, W.; Barrett, F.F.; Devincenzo, J.P. Adenovirus 7a: A community-acquired outbreak in a children’s hospital. Pediatr. Infect. Dis. J. 2000, 19, 996–1000. [Google Scholar] [CrossRef]
  97. Lynch, J.P.; Kajon, A.E. Adenovirus: Epidemiology, Global Spread of Novel Serotypes, and Advances in Treatment and Prevention. Semin. Respir. Crit. Care Med. 2016, 37, 586–602. [Google Scholar] [CrossRef]
  98. Kajon, A.E.; Moseley, J.M.; Metzgar, D.; Huong, H.-S.; Wadleigh, A.; Ryan, M.A.K.; Russell, K.L. Molecular epidemiology of adenovirus type 4 infections in US military recruits in the postvaccination era (1997–2003). J. Infect. Dis. 2007, 196, 67–75. [Google Scholar] [CrossRef] [PubMed]
  99. Ryan, M.A.K.; Gray, G.C.; Smith, B.; McKeehan, J.A.; Hawksworth, A.W.; Malasig, M.D. Large Epidemic of Respiratory Illness Due to Adenovirus Types 7 and 3 in Healthy Young Adults. Clin. Infect. Dis. 2002, 34, 577–582. [Google Scholar] [CrossRef] [PubMed]
  100. Ison, M.G. Adenovirus infections in transplant recipients. Clin. Infect. Dis. 2006, 43, 331–339. [Google Scholar] [CrossRef]
  101. Echavarría, M. Adenoviruses in immunocompromised hosts. Clin. Microbiol. Rev. 2008, 21, 704–715. [Google Scholar] [CrossRef] [PubMed]
  102. Dagan, R.; Schwartz, R.H.; Insel, R.A.; Menegus, M.A. Severe diffuse adenovirus 7a pneumonia in a child with combined immunodeficiency: Possible therapeutic effect of human immune serum globulin containing specific neutralizing antibody. Pediatr. Infect. Dis. 1984, 3, 246–251. [Google Scholar] [CrossRef]
  103. Ferdman, R.M.; Ross, L.; Inderlied, C.; Church, J.A. Adenovirus viremia in human immunodeficiency virus-infected children. Pediatr. Infect. Dis. J. 1997, 16, 413–415. [Google Scholar] [CrossRef]
  104. Ismail, A.M.; Zhou, X.; Dyer, D.W.; Seto, D.; Rajaiya, J.; Chodosh, J. Genomic foundations of evolution and ocular pathogenesis in human adenovirus species D. FEBS Lett. 2019, 593, 3583–3608. [Google Scholar] [CrossRef]
  105. Georgi, F.; Greber, U.F. The Adenovirus Death Protein—A small membrane protein controls cell lysis and disease. FEBS Lett. 2020, 594, 1861–1878. [Google Scholar] [CrossRef]
  106. Zhang, W.; Huang, L. Genome Analysis of A Novel Recombinant Human Adenovirus Type 1 in China. Sci. Rep. 2019, 9, 4298. [Google Scholar] [CrossRef]
  107. Biserni, G.B.; Dondi, A.; Masetti, R.; Bandini, J.; Dormi, A.; Conti, F.; Pession, A.; Lanari, M. Immune Response against Adenovirus in Acute Upper Respiratory Tract Infections in Immunocompetent Children. Vaccines 2020, 8, 602. [Google Scholar] [CrossRef]
  108. Lion, T. Adenovirus infections in immunocompetent and immunocompromised patients. Clin. Microbiol. Rev. 2014, 27, 441–462. [Google Scholar] [CrossRef]
  109. Atasheva, S.; Shayakhmetov, D.M. Cytokine Responses to Adenovirus and Adenovirus Vectors. Viruses 2022, 14, 888. [Google Scholar] [CrossRef]
  110. Kühl, U.; Pauschinger, M.; Seeberg, B.; Lassner, D.; Noutsias, M.; Poller, W.; Schultheiss, H.-P. Viral persistence in the myocardium is associated with progressive cardiac dysfunction. Circulation 2005, 112, 1965–1970. [Google Scholar] [CrossRef] [PubMed]
  111. Proenca-Modena, J.L.; de Souza Cardoso, R.; Criado, M.F.; Milanez, G.P.; de Souza, W.M.; Parise, P.L.; Bertol, J.W.; de Jesus, B.L.S.; Prates, M.C.M.; Silva, M.L.; et al. Human adenovirus replication and persistence in hypertrophic adenoids and palatine tonsils in children. J. Med. Virol. 2019, 91, 1250–1262. [Google Scholar] [CrossRef] [PubMed]
  112. Kosulin, K.; Geiger, E.; Vécsei, A.; Huber, W.-D.; Rauch, M.; Brenner, E.; Wrba, F.; Hammer, K.; Innerhofer, A.; Pötschger, U.; et al. Persistence and reactivation of human adenoviruses in the gastrointestinal tract. Clin. Microbiol. Infect. 2016, 22, 381.e1–381.e8. [Google Scholar] [CrossRef]
  113. Kosulin, K.; Haberler, C.; Hainfellner, J.A.; Amann, G.; Lang, S.; Lion, T. Investigation of adenovirus occurrence in pediatric tumor entities. J. Virol. 2007, 81, 7629–7635. [Google Scholar] [CrossRef]
  114. Kosulin, K.; Rauch, M.; Ambros, P.F.; Pötschger, U.; Chott, A.; Jäger, U.; Drach, J.; Nader, A.; Lion, T. Screening for adenoviruses in haematological neoplasia: High prevalence in mantle cell lymphoma. Eur. J. Cancer 2014, 50, 622–627. [Google Scholar] [CrossRef]
  115. Lion, T. Adenovirus persistence, reactivation, and clinical management. FEBS Lett. 2019, 593, 3571–3582. [Google Scholar] [CrossRef] [PubMed]
  116. Nakai, H.; Yant, S.R.; Storm, T.A.; Fuess, S.; Meuse, L.; Kay, M.A. Extrachromosomal recombinant adeno-associated virus vector genomes are primarily responsible for stable liver transduction in vivo. J. Virol. 2001, 75, 6969–6976. [Google Scholar] [CrossRef]
  117. Kotterman, M.A.; Schaffer, D.V. Engineering adeno-associated viruses for clinical gene therapy. Nat. Rev. Genet. 2014, 15, 445–451. [Google Scholar] [CrossRef]
  118. Duigou, G.J.; Young, C.S.H. Replication-competent adenovirus formation in 293 cells: The recombination-based rate is influenced by structure and location of the transgene cassette and not increased by overproduction of HsRad51, Rad51-interacting, or E2F family proteins. J. Virol. 2005, 79, 5437–5444. [Google Scholar] [CrossRef]
  119. McConnell, M.J.; Imperiale, M.J. Biology of adenovirus and its use as a vector for gene therapy. Hum. Gene Ther. 2004, 15, 1022–1033. [Google Scholar] [CrossRef]
  120. Ewer, K.J.; Lambe, T.; Rollier, C.S.; Spencer, A.J.; Hill, A.V.; Dorrell, L. Viral vectors as vaccine platforms: From immunogenicity to impact. Curr. Opin. Immunol. 2016, 41, 47–54. [Google Scholar] [CrossRef]
  121. Gorziglia, M.I.; Lapcevich, C.; Roy, S.; Kang, Q.; Kadan, M.; Wu, V.; Pechan, P.; Kaleko, M. Generation of an adenovirus vector lacking E1, e2a, E3, and all of E4 except open reading frame 3. J. Virol. 1999, 73, 6048–6055. [Google Scholar] [CrossRef]
  122. Sayedahmed, E.E.; Kumari, R.; Mittal, S.K. Current Use of Adenovirus Vectors and Their Production Methods. Methods Mol. Biol. 2019, 1937, 155–175. [Google Scholar] [CrossRef]
  123. Vemula, S.V.; Mittal, S.K. Production of adenovirus vectors and their use as a delivery system for influenza vaccines. Expert Opin. Biol. Ther. 2010, 10, 1469–1487. [Google Scholar] [CrossRef]
  124. Lee, C.S.; Bishop, E.S.; Zhang, R.; Yu, X.; Farina, E.M.; Yan, S.; Zhao, C.; Zheng, 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]
  125. Brough, D.E.; Lizonova, A.; Hsu, C.; Kulesa, V.A.; Kovesdi, I. A gene transfer vector-cell line system for complete functional complementation of adenovirus early regions E1 and E4. J. Virol. 1996, 70, 6497–6501. [Google Scholar] [CrossRef]
  126. Zhang, X.; Godbey, W.T. Viral vectors for gene delivery in tissue engineering. Adv. Drug Deliv. Rev. 2006, 58, 515–534. [Google Scholar] [CrossRef]
  127. Lusky, M.; Christ, M.; Rittner, K.; Dieterle, A.; Dreyer, D.; Mourot, B.; Schultz, H.; Stoeckel, F.; Pavirani, A.; Mehtali, M. In vitro and in vivo biology of recombinant adenovirus vectors with E1, E1/E2A, or E1/E4 deleted. J. Virol. 1998, 72, 2022–2032. [Google Scholar] [CrossRef]
  128. Mitani, K.; Graham, F.L.; Caskey, C.T.; Kochanek, S. Rescue, propagation, and partial purification of a helper virus-dependent adenovirus vector. Proc. Natl. Acad. Sci. USA 1995, 92, 3854–3858. [Google Scholar] [CrossRef]
  129. Parks, R.J.; Chen, L.; Anton, M.; Sankar, U.; Rudnicki, M.A.; Graham, F.L. A helper-dependent adenovirus vector system: Removal of helper virus by Cre-mediated excision of the viral packaging signal. Proc. Natl. Acad. Sci. USA 1996, 93, 13565–13570. [Google Scholar] [CrossRef]
  130. Palmer, D.J.; Ng, P. Methods for the production of helper-dependent adenoviral vectors. Methods Mol. Biol. 2008, 433, 33–53. [Google Scholar] [CrossRef]
  131. Cervantes-García, D.; Ortiz-López, R.; Mayek-Pérez, N.; Rojas-Martínez, A. Oncolytic virotherapy. Ann. Hepatol. 2008, 7, 34–45. [Google Scholar] [CrossRef]
  132. Gouvarchin Ghaleh, H.E.; Bolandian, M.; Dorostkar, R.; Jafari, A.; Pour, M.F. Concise review on optimized methods in production and transduction of lentiviral vectors in order to facilitate immunotherapy and gene therapy. Biomed. Pharmacother. 2020, 128, 110276. [Google Scholar] [CrossRef]
  133. Available online: https://www.researchamerica.org/blog/ad-ding-value-the-science-behind-adenovirus-vector-vaccines/ (accessed on 1 September 2023).
  134. He, T.C.; Zhou, S.; da Costa, L.T.; Yu, J.; Kinzler, K.W.; Vogelstein, B. A simplified system for generating recombinant adenoviruses. Proc. Natl. Acad. Sci. USA 1998, 95, 2509–2514. [Google Scholar] [CrossRef]
  135. Ng, P.; Parks, R.J.; Cummings, D.T.; Evelegh, C.M.; Sankar, U.; Graham, F.L. A high-efficiency Cre/loxP-based system for construction of adenoviral vectors. Hum. Gene Ther. 1999, 10, 2667–2672. [Google Scholar] [CrossRef]
  136. Syyam, A.; Nawaz, A.; Ijaz, A.; Sajjad, U.; Fazil, A.; Irfan, S.; Muzaffar, A.; Shahid, M.; Idrees, M.; Malik, K.; et al. Adenovirus vector system: Construction, history and therapeutic applications. Biotechniques 2022, 73, 297–305. [Google Scholar] [CrossRef]
  137. Von Seggern, D.J.; Nemerow, G.R. Adenoviral Vectors for Protein Expression. Gene Expr. Syst. 2007, 5, 111–156. [Google Scholar] [CrossRef]
  138. Elahi, S.M.; Jiang, J.; Nazemi-Moghaddam, N.; Gilbert, R. A Method to Generate and Rescue Recombinant Adenovirus Devoid of Replication-Competent Particles in Animal-Origin-Free Culture Medium. Viruses 2023, 15, 2152. [Google Scholar] [CrossRef] [PubMed]
  139. Jalšić, L.; Lytvyn, V.; Elahi, S.M.; Hrapovic, S.; Nassoury, N.; Chahal, P.S.; Gaillet, B.; Gilbert, R. Inducible HEK293 AAV packaging cell lines expressing Rep proteins. Mol. Ther. Methods Clin. Dev. 2023, 30, 259–275. [Google Scholar] [CrossRef]
  140. Graham, F.L.; Smiley, J.; Russell, W.C.; Nairn, R. Characteristics of a human cell line transformed by DNA from human adenovirus type 5. J. Gen. Virol. 1977, 36, 59–74. [Google Scholar] [CrossRef]
  141. Shaw, G.; Morse, S.; Ararat, M.; Graham, F.L. Preferential transformation of human neuronal cells by human adenoviruses and the origin of HEK 293 cells. FASEB J. 2002, 16, 869–871. [Google Scholar] [CrossRef]
  142. Fallaux, F.J.; Kranenburg, O.; Cramer, S.J.; Houweling, A.; van Ormondt, H.; Hoeben, R.C.; van der Eb, A.J. Characterization of 911: A new helper cell line for the titration and propagation of early region 1-deleted adenoviral vectors. Hum. Gene Ther. 1996, 7, 215–222. [Google Scholar] [CrossRef]
  143. Imler, J.L.; Chartier, C.; Dreyer, D.; Dieterle, A.; Sainte-Marie, M.; Faure, T.; Pavirani, A.; Mehtali, M. Novel complementation cell lines derived from human lung carcinoma A549 cells support the growth of E1-deleted adenovirus vectors. Gene Ther. 1996, 3, 75–84. [Google Scholar]
  144. Murakami, P.; Pungor, E.; Files, J.; Do, L.; van Rijnsoever, R.; Vogels, R.; Bout, A.; McCaman, M. A single short stretch of homology between adenoviral vector and packaging cell line can give rise to cytopathic effect-inducing, helper-dependent E1-positive particles. Hum. Gene Ther. 2002, 13, 909–920. [Google Scholar] [CrossRef]
  145. Krougliak, V.; Graham, F.L. Development of cell lines capable of complementing E1, E4, and protein IX defective adenovirus type 5 mutants. Hum. Gene Ther. 1995, 6, 1575–1586. [Google Scholar] [CrossRef]
  146. Zhou, H.; Beaudet, A.L. A new vector system with inducible E2a cell line for production of higher titer and safer adenoviral vectors. Virology 2000, 275, 348–357. [Google Scholar] [CrossRef]
  147. Klessig, D.F.; Grodzicker, T.; Cleghon, V. Construction of human cell lines which contain and express the adenovirus DNA binding protein gene by cotransformation with the HSV-1 tk gene. Virus Res. 1984, 1, 169–188. [Google Scholar] [CrossRef]
  148. Zhou, H.; O’Neal, W.; Morral, N.; Beaudet, A.L. Development of a complementing cell line and a system for construction of adenovirus vectors with E1 and E2a deleted. J. Virol. 1996, 70, 7030–7038. [Google Scholar] [CrossRef]
  149. Segura, M.M.; Alba, R.; Bosch, A.; Chillón, M. Advances in helper-dependent adenoviral vector research. Curr. Gene Ther. 2008, 8, 222–235. [Google Scholar] [CrossRef]
  150. Hardy, S.; Kitamura, M.; Harris-Stansil, T.; Dai, Y.; Phipps, M.L. Construction of adenovirus vectors through Cre-lox recombination. J. Virol. 1997, 71, 1842–1849. [Google Scholar] [CrossRef]
  151. Kovesdi, I.; Hedley, S.J. Adenoviral producer cells. Viruses 2010, 2, 1681–1703. [Google Scholar] [CrossRef] [PubMed]
  152. Tan, R.; Li, C.; Jiang, S.; Ma, L. A novel and simple method for construction of recombinant adenoviruses. Nucleic Acids Res. 2006, 34, e89. [Google Scholar] [CrossRef]
  153. Su, Q.; Sena-Esteves, M.; Gao, G. Purification of the Recombinant Adenovirus by Cesium Chloride Gradient Centrifugation. Cold Spring Harb. Protoc. 2019, 5, 374–378. [Google Scholar] [CrossRef]
  154. Stepanenko, A.A.; Sosnovtseva, A.O.; Valikhov, M.P.; Chekhonin, V.P. A new insight into aggregation of oncolytic adenovirus Ad5-delta-24-RGD during CsCl gradient ultracentrifugation. Sci. Rep. 2021, 11, 16088. [Google Scholar] [CrossRef]
  155. Ferreira, R.G.; Gordon, N.F.; Stock, R.; Petrides, D. Adenoviral Vector COVID-19 Vaccines: Process and Cost Analysis. Processes 2021, 9, 1430. [Google Scholar] [CrossRef]
  156. Travieso, T.; Li, J.; Mahesh, S.; Da Mello, J.F.R.E.; Blasi, M. The use of viral vectors in vaccine development. NPJ Vaccines 2022, 7, 75. [Google Scholar] [CrossRef]
  157. Kajon, A.E.; Lamson, D.M.; St George, K. Emergence and re-emergence of respiratory adenoviruses in the United States. Curr. Opin. Virol. 2019, 34, 63–69. [Google Scholar] [CrossRef]
  158. Crystal, R.G. Adenovirus: The first effective in vivo gene delivery vector. Hum. Gene Ther. 2014, 25, 3–11. [Google Scholar] [CrossRef] [PubMed]
  159. Rosenfeld, M.A.; Siegfried, W.; Yoshimura, K.; Yoneyama, K.; Fukayama, M.; Stier, L.E.; Pääkkö, P.K.; Gilardi, P.; Stratford-Perricaudet, L.D.; Perricaudet, M. Adenovirus-mediated transfer of a recombinant alpha 1-antitrypsin gene to the lung epithelium in vivo. Science 1991, 252, 431–434. [Google Scholar] [CrossRef] [PubMed]
  160. Greenwood, H.L.; Thorsteinsdottir, H.; Perry, G.; Renihan, J.; Singer, P.A.; Daar, A.S. Regenerative medicine: New opportunities for developing countries. IJBT 2006, 8, 60. [Google Scholar] [CrossRef]
  161. Jacques, E.; Suuronen, E.J. The Progression of Regenerative Medicine and its Impact on Therapy Translation. Clin. Transl. Sci. 2020, 13, 440–450. [Google Scholar] [CrossRef] [PubMed]
  162. Hosseinkhani, H.; Domb, A.J.; Sharifzadeh, G.; Nahum, V. Gene Therapy for Regenerative Medicine. Pharmaceutics 2023, 15, 856. [Google Scholar] [CrossRef]
  163. Bukharova, T.B.; Nedorubova, I.A.; Mokrousova, V.O.; Meglei, A.Y.; Basina, V.P.; Nedorubov, A.A.; Vasilyev, A.V.; Grigoriev, T.E.; Zagoskin, Y.D.; Chvalun, S.N.; et al. Adenovirus-Based Gene Therapy for Bone Regeneration: A Comparative Analysis of In Vivo and Ex Vivo BMP2 Gene Delivery. Cells 2023, 12, 1762. [Google Scholar] [CrossRef] [PubMed]
  164. Mack, C.A.; Patel, S.R.; Schwarz, E.A.; Zanzonico, P.; Hahn, R.T.; Ilercil, A.; Devereux, R.B.; Goldsmith, S.J.; Christian, T.F.; Sanborn, T.A.; et al. Biologic bypass with the use of adenovirus-mediated gene transfer of the complementary deoxyribonucleic acid for vascular endothelial growth factor 121 improves myocardial perfusion and function in the ischemic porcine heart. J. Thorac. Cardiovasc. Surg. 1998, 115, 168–176; discussion 176–177. [Google Scholar] [CrossRef]
  165. Schwarz, E.R.; Speakman, M.T.; Patterson, M.; Hale, S.S.; Isner, J.M.; Kedes, L.H.; Kloner, R.A. Evaluation of the effects of intramyocardial injection of DNA expressing vascular endothelial growth factor (VEGF) in a myocardial infarction model in the rat--angiogenesis and angioma formation. J. Am. Coll. Cardiol. 2000, 35, 1323–1330. [Google Scholar] [CrossRef] [PubMed]
  166. Pajula, J.; Lähteenvuo, J.; Lähteenvuo, M.; Honkonen, K.; Halonen, P.; Hätinen, O.-P.; Kuivanen, A.; Heikkilä, M.; Nurro, J.; Hartikainen, J.; et al. Adenoviral VEGF-DΔN ΔC gene therapy for myocardial ischemia. Front. Bioeng. Biotechnol. 2022, 10, 999226. [Google Scholar] [CrossRef] [PubMed]
  167. McCann, N.; O’Connor, D.; Lambe, T.; Pollard, A.J. Viral vector vaccines. Curr. Opin. Immunol. 2022, 77, 102210. [Google Scholar] [CrossRef] [PubMed]
  168. Appledorn, D.M.; Patial, S.; McBride, A.; Godbehere, S.; van Rooijen, N.; Parameswaran, N.; Amalfitano, A. Adenovirus vector-induced innate inflammatory mediators, MAPK signaling, as well as adaptive immune responses are dependent upon both TLR2 and TLR9 in vivo. J. Immunol. 2008, 181, 2134–2144. [Google Scholar] [CrossRef] [PubMed]
  169. Chéneau, C.; Eichholz, K.; Tran, T.H.; Tran, T.T.P.; Paris, O.; Henriquet, C.; Bajramovic, J.J.; Pugniere, M.; Kremer, E.J. Lactoferrin Retargets Human Adenoviruses to TLR4 to Induce an Abortive NLRP3-Associated Pyroptotic Response in Human Phagocytes. Front. Immunol. 2021, 12, 685218. [Google Scholar] [CrossRef] [PubMed]
  170. Rhee, E.G.; Blattman, J.N.; Kasturi, S.P.; Kelley, R.P.; Kaufman, D.R.; Lynch, D.M.; La Porte, A.; Simmons, N.L.; Clark, S.L.; Pulendran, B.; et al. Multiple innate immune pathways contribute to the immunogenicity of recombinant adenovirus vaccine vectors. J. Virol. 2011, 85, 315–323. [Google Scholar] [CrossRef] [PubMed]
  171. Atasheva, S.; Yao, J.; Shayakhmetov, D.M. Innate immunity to adenovirus: Lessons from mice. FEBS Lett. 2019, 593, 3461–3483. [Google Scholar] [CrossRef]
  172. Sakurai, F.; Tachibana, M.; Mizuguchi, H. Adenovirus vector-based vaccine for infectious diseases. Drug Metab. Pharmacokinet. 2022, 42, 100432. [Google Scholar] [CrossRef]
  173. Elkashif, A.; Alhashimi, M.; Sayedahmed, E.E.; Sambhara, S.; Mittal, S.K. Adenoviral vector-based platforms for developing effective vaccines to combat respiratory viral infections. Clin. Transl. Immunol. 2021, 10, e1345. [Google Scholar] [CrossRef]
  174. Marquez-Martinez, S.; Vijayan, A.; Khan, S.; Zahn, R. Cell entry and innate sensing shape adaptive immune responses to adenovirus-based vaccines. Curr. Opin. Immunol. 2023, 80, 102282. [Google Scholar] [CrossRef]
  175. Tatsis, N.; Ertl, H.C.J. Adenoviruses as vaccine vectors. Mol. Ther. 2004, 10, 616–629. [Google Scholar] [CrossRef]
  176. Baden, L.R.; Liu, J.; Li, H.; Johnson, J.A.; Walsh, S.R.; Kleinjan, J.A.; Engelson, B.A.; Peter, L.; Abbink, P.; Milner, D.A.; et al. Induction of HIV-1-specific mucosal immune responses following intramuscular recombinant adenovirus serotype 26 HIV-1 vaccination of humans. J. Infect. Dis. 2015, 211, 518–528. [Google Scholar] [CrossRef]
  177. Provine, N.M.; Amini, A.; Garner, L.C.; Spencer, A.J.; Dold, C.; Hutchings, C.; Silva Reyes, L.; FitzPatrick, M.E.B.; Chinnakannan, S.; Oguti, B.; et al. MAIT cell activation augments adenovirus vector vaccine immunogenicity. Science 2021, 371, 521–526. [Google Scholar] [CrossRef]
  178. Abbink, P.; Lemckert, A.A.C.; Ewald, B.A.; Lynch, D.M.; Denholtz, M.; Smits, S.; Holterman, L.; Damen, I.; Vogels, R.; Thorner, A.R.; et al. Comparative seroprevalence and immunogenicity of six rare serotype recombinant adenovirus vaccine vectors from subgroups B and D. J. Virol. 2007, 81, 4654–4663. [Google Scholar] [CrossRef]
  179. Holterman, L.; Vogels, R.; van der Vlugt, R.; Sieuwerts, M.; Grimbergen, J.; Kaspers, J.; Geelen, E.; van der Helm, E.; Lemckert, A.; Gillissen, G.; et al. Novel replication-incompetent vector derived from adenovirus type 11 (Ad11) for vaccination and gene therapy: Low seroprevalence and non-cross-reactivity with Ad5. J. Virol. 2004, 78, 13207–13215. [Google Scholar] [CrossRef]
  180. Roberts, D.M.; Nanda, A.; Havenga, M.J.E.; Abbink, P.; Lynch, D.M.; Ewald, B.A.; Liu, J.; Thorner, A.R.; Swanson, P.E.; Gorgone, D.A.; et al. Hexon-chimaeric adenovirus serotype 5 vectors circumvent pre-existing anti-vector immunity. Nature 2006, 441, 239–243. [Google Scholar] [CrossRef]
  181. Barouch, D.H.; Liu, J.; Lynch, D.M.; O’Brien, K.L.; La Porte, A.; Simmons, N.L.; Riggs, A.M.; Clark, S.; Abbink, P.; Montefiori, D.C.; et al. Protective efficacy of a single immunization of a chimeric adenovirus vector-based vaccine against simian immunodeficiency virus challenge in rhesus monkeys. J. Virol. 2009, 83, 9584–9590. [Google Scholar] [CrossRef]
  182. Farina, S.F.; Gao, G.P.; Xiang, Z.Q.; Rux, J.J.; Burnett, R.M.; Alvira, M.R.; Marsh, J.; Ertl, H.C.; Wilson, J.M. Replication-defective vector based on a chimpanzee adenovirus. J. Virol. 2001, 75, 11603–11613. [Google Scholar] [CrossRef]
  183. O’Riordan, C.R.; Lachapelle, A.; Delgado, C.; Parkes, V.; Wadsworth, S.C.; Smith, A.E.; Francis, G.E. PEGylation of adenovirus with retention of infectivity and protection from neutralizing antibody in vitro and in vivo. Hum. Gene Ther. 1999, 10, 1349–1358. [Google Scholar] [CrossRef]
  184. Yotnda, P.; Chen, D.-H.; Chiu, W.; Piedra, P.A.; Davis, A.; Templeton, N.S.; Brenner, M.K. Bilamellar cationic liposomes protect adenovectors from preexisting humoral immune responses. Mol. Ther. 2002, 5, 233–241. [Google Scholar] [CrossRef] [PubMed]
  185. Zhang, Y.; Wu, J.; Zhang, H.; Wei, J.; Wu, J. Extracellular Vesicles-Mimetic Encapsulation Improves Oncolytic Viro-Immunotherapy in Tumors With Low Coxsackie and Adenovirus Receptor. Front. Bioeng. Biotechnol. 2020, 8, 574007. [Google Scholar] [CrossRef]
  186. Kremer, E.J. Pros and Cons of Adenovirus-Based SARS-CoV-2 Vaccines. Mol. Ther. 2020, 28, 2303–2304. [Google Scholar] [CrossRef]
  187. Hasanpourghadi, M.; Novikov, M.; Ertl, H.C.J. COVID-19 Vaccines Based on Adenovirus Vectors. Trends Biochem. Sci. 2021, 46, 429–430. [Google Scholar] [CrossRef]
  188. Tumban, E. Lead SARS-CoV-2 Candidate Vaccines: Expectations from Phase III Trials and Recommendations Post-Vaccine Approval. Viruses 2020, 13, 54. [Google Scholar] [CrossRef]
  189. Voysey, M.; Clemens, S.A.C.; Madhi, S.A.; Weckx, L.Y.; Folegatti, P.M.; Aley, P.K.; Angus, B.; Baillie, V.L.; Barnabas, S.L.; Bhorat, Q.E.; et al. Safety and efficacy of the ChAdOx1 nCoV-19 vaccine (AZD1222) against SARS-CoV-2: An interim analysis of four randomised controlled trials in Brazil, South Africa, and the UK. Lancet 2021, 397, 99–111. [Google Scholar] [CrossRef]
  190. Jones, I.; Roy, P. Sputnik V COVID-19 vaccine candidate appears safe and effective. Lancet 2021, 397, 642–643. [Google Scholar] [CrossRef]
  191. Huang, Z.; Xu, S.; Liu, J.; Wu, L.; Qiu, J.; Wang, N.; Ren, J.; Li, Z.; Guo, X.; Tao, F.; et al. Effectiveness of inactivated and Ad5-nCoV COVID-19 vaccines against SARS-CoV-2 Omicron BA. 2 variant infection, severe illness, and death. BMC Med. 2022, 20, 400. [Google Scholar] [CrossRef]
  192. Alter, G.; Yu, J.; Liu, J.; Chandrashekar, A.; Borducchi, E.N.; Tostanoski, L.H.; McMahan, K.; Jacob-Dolan, C.; Martinez, D.R.; Chang, A.; et al. Immunogenicity of Ad26.COV2.S vaccine against SARS-CoV-2 variants in humans. Nature 2021, 596, 268–272. [Google Scholar] [CrossRef]
  193. Stanley, D.A.; Honko, A.N.; Asiedu, C.; Trefry, J.C.; Lau-Kilby, A.W.; Johnson, J.C.; Hensley, L.; Ammendola, V.; Abbate, A.; Grazioli, F.; et al. Chimpanzee adenovirus vaccine generates acute and durable protective immunity against ebolavirus challenge. Nat. Med. 2014, 20, 1126–1129. [Google Scholar] [CrossRef] [PubMed]
  194. Salisch, N.C.; Stephenson, K.E.; Williams, K.; Cox, F.; van der Fits, L.; Heerwegh, D.; Truyers, C.; Habets, M.N.; Kanjilal, D.G.; Larocca, R.A.; et al. A Double-Blind, Randomized, Placebo-Controlled Phase 1 Study of Ad26.ZIKV.001, an Ad26-Vectored Anti-Zika Virus Vaccine. Ann. Intern. Med. 2021, 174, 585–594. [Google Scholar] [CrossRef] [PubMed]
  195. Gray, G.; Buchbinder, S.; Duerr, A. Overview of STEP and Phambili trial results: Two phase IIb test-of-concept studies investigating the efficacy of MRK adenovirus type 5 gag/pol/nef subtype B HIV vaccine. Curr. Opin. HIV AIDS 2010, 5, 357–361. [Google Scholar] [CrossRef] [PubMed]
  196. Baden, L.R.; Stieh, D.J.; Sarnecki, M.; Walsh, S.R.; Tomaras, G.D.; Kublin, J.G.; McElrath, M.J.; Alter, G.; Ferrari, G.; Montefiori, D.; et al. Safety and immunogenicity of two heterologous HIV vaccine regimens in healthy, HIV-uninfected adults (TRAVERSE): A randomised, parallel-group, placebo-controlled, double-blind, phase 1/2a study. Lancet HIV 2020, 7, e688–e698. [Google Scholar] [CrossRef]
  197. van Kampen, K.R.; Shi, Z.; Gao, P.; Zhang, J.; Foster, K.W.; Chen, D.-T.; Marks, D.; Elmets, C.A.; Tang, D.C. Safety and immunogenicity of adenovirus-vectored nasal and epicutaneous influenza vaccines in humans. Vaccine 2005, 23, 1029–1036. [Google Scholar] [CrossRef]
  198. Khalil, A.M. The genome editing revolution: Review. J. Genet. Eng. Biotechnol. 2020, 18, 68. [Google Scholar] [CrossRef]
  199. Redman, M.; King, A.; Watson, C.; King, D. What is CRISPR/Cas9? Arch. Dis. Child. Educ. Pract. Ed. 2016, 101, 213–215. [Google Scholar] [CrossRef]
  200. Naso, M.F.; Tomkowicz, B.; Perry, W.L.; Strohl, W.R. Adeno-Associated Virus (AAV) as a Vector for Gene Therapy. BioDrugs 2017, 31, 317–334. [Google Scholar] [CrossRef] [PubMed]
  201. Maggio, I.; Liu, J.; Janssen, J.M.; Chen, X.; Gonçalves, M.A.F.V. Adenoviral vectors encoding CRISPR/Cas9 multiplexes rescue dystrophin synthesis in unselected populations of DMD muscle cells. Sci. Rep. 2016, 6, 37051. [Google Scholar] [CrossRef]
  202. Stephens, C.J.; Kashentseva, E.; Everett, W.; Kaliberova, L.; Curiel, D.T. Targeted in vivo knock-in of human alpha-1-antitrypsin cDNA using adenoviral delivery of CRISPR/Cas9. Gene Ther. 2018, 25, 139–156. [Google Scholar] [CrossRef] [PubMed]
  203. Tsukamoto, T.; Sakai, E.; Iizuka, S.; Taracena-Gándara, M.; Sakurai, F.; Mizuguchi, H. Generation of the Adenovirus Vector-Mediated CRISPR/Cpf1 System and the Application for Primary Human Hepatocytes Prepared from Humanized Mice with Chimeric Liver. Biol. Pharm. Bull. 2018, 41, 1089–1095. [Google Scholar] [CrossRef]
  204. Gao, J.; Zhang, W.; Ehrhardt, A. Expanding the Spectrum of Adenoviral Vectors for Cancer Therapy. Cancers 2020, 12, 1139. [Google Scholar] [CrossRef]
  205. Mulvihill, S.; Warren, R.; Venook, A.; Adler, A.; Randlev, B.; Heise, C.; Kirn, D. Safety and feasibility of injection with an E1B-55 kDa gene-deleted, replication-selective adenovirus (ONYX-015) into primary carcinomas of the pancreas: A phase I trial. Gene Ther. 2001, 8, 308–315. [Google Scholar] [CrossRef] [PubMed]
  206. Gryciuk, A.; Rogalska, M.; Baran, J.; Kuryk, L.; Staniszewska, M. Oncolytic Adenoviruses Armed with Co-Stimulatory Molecules for Cancer Treatment. Cancers 2023, 15, 1947. [Google Scholar] [CrossRef] [PubMed]
  207. Bischoff, J.R.; Kirn, D.H.; Williams, A.; Heise, C.; Horn, S.; Muna, M.; Ng, L.; Nye, J.A.; Sampson-Johannes, A.; Fattaey, A.; et al. An adenovirus mutant that replicates selectively in p53-deficient human tumor cells. Science 1996, 274, 373–376. [Google Scholar] [CrossRef]
  208. Heise, C.; Hermiston, T.; Johnson, L.; Brooks, G.; Sampson-Johannes, A.; Williams, A.; Hawkins, L.; Kirn, D. An adenovirus E1A mutant that demonstrates potent and selective systemic anti-tumoral efficacy. Nat. Med. 2000, 6, 1134–1139. [Google Scholar] [CrossRef]
  209. Kitajima, S.; Li, F.; Takahashi, C. Tumor Milieu Controlled by RB Tumor Suppressor. Int. J. Mol. Sci. 2020, 21, 2450. [Google Scholar] [CrossRef]
  210. Lang, F.F.; Conrad, C.; Gomez-Manzano, C.; Yung, W.K.A.; Sawaya, R.; Weinberg, J.S.; Prabhu, S.S.; Rao, G.; Fuller, G.N.; Aldape, K.D.; et al. Phase I Study of DNX-2401 (Delta-24-RGD) Oncolytic Adenovirus: Replication and Immunotherapeutic Effects in Recurrent Malignant Glioma. J. Clin. Oncol. 2018, 36, 1419–1427. [Google Scholar] [CrossRef]
  211. Mach, N.; Gao, J.; Schaffarczyk, L.; Janz, S.; Ehrke-Schulz, E.; Dittmar, T.; Ehrhardt, A.; Zhang, W. Spectrum-Wide Exploration of Human Adenoviruses for Breast Cancer Therapy. Cancers 2020, 12, 1403. [Google Scholar] [CrossRef] [PubMed]
  212. Chen, C.Y.; Senac, J.S.; Weaver, E.A.; May, S.M.; Jelinek, D.F.; Greipp, P.; Witzig, T.; Barry, M.A. Species D adenoviruses as oncolytics against B-cell cancers. Clin. Cancer Res. 2011, 17, 6712–6722. [Google Scholar] [CrossRef]
  213. Persson, B.D.; John, L.; Rafie, K.; Strebl, M.; Frängsmyr, L.; Ballmann, M.Z.; Mindler, K.; Havenga, M.; Lemckert, A.; Stehle, T.; et al. Human species D adenovirus hexon capsid protein mediates cell entry through a direct interaction with CD46. Proc. Natl. Acad. Sci. USA 2021, 118, e2020732118. [Google Scholar] [CrossRef]
  214. Daussy, C.F.; Pied, N.; Wodrich, H. Understanding Post Entry Sorting of Adenovirus Capsids; A Chance to Change Vaccine Vector Properties. Viruses 2021, 13, 1221. [Google Scholar] [CrossRef]
  215. Ranki, T.; Pesonen, S.; Hemminki, A.; Partanen, K.; Kairemo, K.; Alanko, T.; Lundin, J.; Linder, N.; Turkki, R.; Ristimäki, A.; et al. Phase I study with ONCOS-102 for the treatment of solid tumors—An evaluation of clinical response and exploratory analyses of immune markers. J. Immunother. Cancer 2016, 4, 17. [Google Scholar] [CrossRef] [PubMed]
  216. Bett, A.J.; Haddara, W.; Prevec, L.; Graham, F.L. An efficient and flexible system for construction of adenovirus vectors with insertions or deletions in early regions 1 and 3. Proc. Natl. Acad. Sci. USA 1994, 91, 8802–8806. [Google Scholar] [CrossRef] [PubMed]
  217. Mantwill, K.; Klein, F.G.; Wang, D.; Hindupur, S.V.; Ehrenfeld, M.; Holm, P.S.; Nawroth, R. Concepts in Oncolytic Adenovirus Therapy. Int. J. Mol. Sci. 2021, 22, 10522. [Google Scholar] [CrossRef] [PubMed]
  218. Blackburn, R.V.; Galoforo, S.S.; Corry, P.M.; Lee, Y.J. Adenoviral-mediated transfer of a heat-inducible double suicide gene into prostate carcinoma cells. Cancer Res. 1998, 58, 1358–1362. [Google Scholar] [PubMed]
  219. Robinson, M.; Ge, Y.; Ko, D.; Yendluri, S.; Laflamme, G.; Hawkins, L.; Jooss, K. Comparison of the E3 and L3 regions for arming oncolytic adenoviruses to achieve a high level of tumor-specific transgene expression. Cancer Gene Ther. 2008, 15, 9–17. [Google Scholar] [CrossRef]
  220. Shoushtari, A.N.; Olszanski, A.J.; Nyakas, M.; Hornyak, T.J.; Wolchok, J.D.; Levitsky, V.; Kuryk, L.; Hansen, T.B.; Jäderberg, M. Pilot Study of ONCOS-102 and Pembrolizumab: Remodeling of the Tumor Microenvironment and Clinical Outcomes in Anti-PD-1-Resistant Advanced Melanoma. Clin. Cancer Res. 2023, 29, 100–109. [Google Scholar] [CrossRef]
  221. Havunen, R.; Kalliokoski, R.; Siurala, M.; Sorsa, S.; Santos, J.M.; Cervera-Carrascon, V.; Anttila, M.; Hemminki, A. Cytokine-Coding Oncolytic Adenovirus TILT-123 Is Safe, Selective, and Effective as a Single Agent and in Combination with Immune Checkpoint Inhibitor Anti-PD-1. Cells 2021, 10, 246. [Google Scholar] [CrossRef]
  222. Zhang, Q.; Zhang, J.; Tian, Y.; Zhu, G.; Liu, S.; Liu, F. Efficacy of a novel double-controlled oncolytic adenovirus driven by the Ki67 core promoter and armed with IL-15 against glioblastoma cells. Cell Biosci. 2020, 10, 124. [Google Scholar] [CrossRef]
  223. Salzwedel, A.O.; Han, J.; LaRocca, C.J.; Shanley, R.; Yamamoto, M.; Davydova, J. Combination of interferon-expressing oncolytic adenovirus with chemotherapy and radiation is highly synergistic in hamster model of pancreatic cancer. Oncotarget 2018, 9, 18041–18052. [Google Scholar] [CrossRef]
  224. Cervera-Carrascon, V.; Siurala, M.; Santos, J.M.; Havunen, R.; Tähtinen, S.; Karell, P.; Sorsa, S.; Kanerva, A.; Hemminki, A. TNFa and IL-2 armed adenoviruses enable complete responses by anti-PD-1 checkpoint blockade. Oncoimmunology 2018, 7, e1412902. [Google Scholar] [CrossRef]
  225. Wang, G.; Zhang, Z.; Zhong, K.; Wang, Z.; Yang, N.; Tang, X.; Li, H.; Lu, Q.; Wu, Z.; Yuan, B.; et al. CXCL11-armed oncolytic adenoviruses enhance CAR-T cell therapeutic efficacy and reprogram tumor microenvironment in glioblastoma. Mol. Ther. 2023, 31, 134–153. [Google Scholar] [CrossRef] [PubMed]
  226. Quixabeira, D.C.A.; Pakola, S.; Jirovec, E.; Havunen, R.; Basnet, S.; Santos, J.M.; Kudling, T.V.; Clubb, J.H.A.; Haybout, L.; Arias, V.; et al. Boosting cytotoxicity of adoptive allogeneic NK cell therapy with an oncolytic adenovirus encoding a human vIL-2 cytokine for the treatment of human ovarian cancer. Cancer Gene Ther. 2023, 30, 1679–1690. [Google Scholar] [CrossRef] [PubMed]
  227. Su, Y.; Li, J.; Ji, W.; Wang, G.; Fang, L.; Zhang, Q.; Ang, L.; Zhao, M.; Sen, Y.; Chen, L.; et al. Triple-serotype chimeric oncolytic adenovirus exerts multiple synergistic mechanisms against solid tumors. J. Immunother. Cancer 2022, 10, e004691. [Google Scholar] [CrossRef]
Table 1. Application of an adenoviral vector in a vaccine clinical trial.
Table 1. Application of an adenoviral vector in a vaccine clinical trial.
DiseasePhaseClinical Trial Number
HIVINCT00479999
EbolaINCT02289027
HIVINCT01989533
MalariaINCT00371189
COVID-19 diseaseIINCT05027672
COVID-19 diseaseINCT04568811
Prostata cancerIINCT00583752
TuberculosisINCT00800670
Hepatitis CI NCT01094873
MelanomaII NCT00010309
Table 2. Some selected clinical trials of oncolytic adenoviral vectors for anti-cancer therapy recruiting in the year 2023.
Table 2. Some selected clinical trials of oncolytic adenoviral vectors for anti-cancer therapy recruiting in the year 2023.
OAdTransgeneDiseasePhaseClinical Trial Number
Ad5-yCD/mut
TKSR39rep-
hIL-12
yCD, TK and hIL-12Prostate adenocarcinomaINCT02555397
Ad11p/Ad3NoSolid epithelial tumorsINCT02636036
VCN-01/Ad5 based)NoSolid tumorsINCT02045602
AdAPT-001TGF-ßSolid tumorsINCT04673942
Ad5/3-E2F-D24-hTNFa-IRES-hIL2 (TILT-123)TNF-α,
IL-2
Solid tumorsINCT04695327
OAdTILT-123NoMelanomaINCT04217473
Ad5-yCD/mut
TKSR39rep-ADP
AstrocytomaINCT05686798
VCN-01 NoRetinoblastomaINCT03284268
Ad5-yCD/mutTKSR39rep-ADPNoPancreas AdenocarcinomaI NCT04739046
NG-641FAP, CXCL9/
CXCL10/IFNa2
Solid epithelial tumorsI NCT04053283
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Scarsella, L.; Ehrke-Schulz, E.; Paulussen, M.; Thal, S.C.; Ehrhardt, A.; Aydin, M. Advances of Recombinant Adenoviral Vectors in Preclinical and Clinical Applications. Viruses 2024, 16, 377. https://doi.org/10.3390/v16030377

AMA Style

Scarsella L, Ehrke-Schulz E, Paulussen M, Thal SC, Ehrhardt A, Aydin M. Advances of Recombinant Adenoviral Vectors in Preclinical and Clinical Applications. Viruses. 2024; 16(3):377. https://doi.org/10.3390/v16030377

Chicago/Turabian Style

Scarsella, Luca, Eric Ehrke-Schulz, Michael Paulussen, Serge C. Thal, Anja Ehrhardt, and Malik Aydin. 2024. "Advances of Recombinant Adenoviral Vectors in Preclinical and Clinical Applications" Viruses 16, no. 3: 377. https://doi.org/10.3390/v16030377

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